paper_id
stringlengths
9
16
version
stringclasses
26 values
yymm
stringclasses
311 values
created
timestamp[s]
title
stringlengths
6
335
secondary_subfield
sequencelengths
1
8
abstract
stringlengths
25
3.93k
primary_subfield
stringclasses
124 values
field
stringclasses
20 values
fulltext
stringlengths
0
2.84M
1312.3238
1
1312
2013-12-11T16:58:15
Unzipping of DNA under the influence of external fields
[ "physics.bio-ph", "cond-mat.soft", "q-bio.BM" ]
We study the features on the unzipping process of a modified version of the Peyrard-Bishop- Dauxois model. We show that the inclusion of a barrier in the on-site potential allows for the existence of stable domain wall solutions between open and closed regions of the DNA chain. We analyze the linear stability of such solutions and study their relevance in the dynamical behavior of DNA under ac forces.
physics.bio-ph
physics
Unzipping of DNA under the influence of external fields A. E. Bergues-Pupo,1, 2 J. M. Bergues,3 and F. Falo2, 4, ∗ 1Departamento de F´ısica, Universidad de Oriente, 90500 Santiago de Cuba, Cuba 2Dpto. de F´ısica de la Materia Condensada, Universidad de Zaragoza. 50009 Zaragoza, Spain 3Escuela Polit´ecnica Superior y Facultad de Ciencias de la Salud, Universidad San Jorge, 50830 Villanueva de G´allego, Zaragoza, Spain 4Instituto de Biocomputaci´on y F´ısica de Sistemas Complejos (BIFI), Universidad de Zaragoza, 50009 Zaragoza, Spain (Dated: November 8, 2018) Abstract We study the features on the unzipping process of a modified version of the Peyrard-Bishop- Dauxois model. We show that the inclusion of a barrier in the on-site potential allows for the existence of stable domain wall solutions between open and closed regions of the DNA chain. We analyze the linear stability of such solutions and study their relevance in the dynamical behavior of DNA under ac forces. PACS numbers: 87.15.-v, 36.20.-r, 87.18.Tt, 83.10.Rs, 05.40.-a Keywords: DNA mechanical unzipping, DNA modeling, linear stability, ac fields 3 1 0 2 c e D 1 1 ] h p - o i b . s c i s y h p [ 1 v 8 3 2 3 . 2 1 3 1 : v i X r a ∗Electronic address: [email protected] 1 I. INTRODUCTION The study of the energy landscape of biomolecules is possible, among other methods, through their mechanical response under the action of an external force. These studies are important because, in many cellular process like transcription or replication, DNA is under the action of different mechanical stresses. One of the experiments that shows these potentialities is the mechanical unzipping of DNA [1 -- 3]. It consists of separating the DNA molecule double helix by pulling apart one strand from the other one by using an atomic force microscope, optical or magnetic tweezers. The obtained signal can give information of base pair (bp) content of a sequence [1 -- 3] or may be used to estimate interaction energies of bps through the chain [4]. Special attention has been given in the last years to the influence of external ac fields on the dynamic response of DNA [5 -- 11]. From one side, the possible effects of THz fields have been discussed [5 -- 8]. The argument for such an influence resides on the fact that weak bonding energies of the molecule, e.g. hydrogen bonds between nucleotides of a bp, are in the THz frequency range and thus resonant effects could be expected. On the other side, ac mechanical forces acting during single molecule experiments have been addressed [9, 10]. In [9] the use of a periodic driving protocol is proposed for the folding and unfolding of a protein and it is obtained a better resolution in the reconstruction of the free energy landscape of the molecule. On the other hand, stochastic resonance and resonant activation were observed during the folding unfolding of short DNA hairpins [10]. The resonant frequencies obtained in this case match the hopping dynamics of the system rather than the natural frequencies and may be used to estimate kinetics rates. The idea of the use of mechanical alternate forces is also supported by the fact that many process occur in a periodic way due to the periodic energy consumption inside the cell [11]. The use of simple physical models may help in the understanding of such kind of processes. In fact, some of the mechanical and thermal properties of DNA, and other biomolecules, can be modeled at a mesoscopic level. One of the most studied models at this scale is the Peyrard-Bishop-Dauxois (PBD) model [12]. The only degree of freedom relevant for this model is the inter-strand separation. Despite its simplicity, it contains the main ingredients to describe the phenomena mentioned above. It was initially proposed for explaining the melting transition of DNA, i.e., the separation of the double strand upon heating, but 2 it has been also used to explain other phenomena as DNA mechanical unzipping [13 -- 16], microscopic mechanical flexibility of DNA [17] and the influence of external ac fields on the DNA dynamics [5 -- 7]. In a recent work [7], we studied the influence of an ac external field on the melting transition and bubble formation of homogeneous and heterogeneous DNA sequences. We used a modified PBD model that includes a solvation barrier that takes into account the interaction of the molecule, once opened, with the solvent. This term has been used in previous works in order to provide results like bubble lifetime and melting width closer to those of experimental data [18 -- 20] and to model the unzipping process under different salt concentrations of the solution [16]. We found that the external field influences resonantly the DNA dynamics at certain frequency values, corresponding approximately to the natural oscillation frequencies of AT and CG bps inside the Morse potential. Consequently the response of the system distinguishes the AT-rich regions from CG-rich regions. Interestingly, if the barrier is not included, this influence is almost independent of the frequency. The introduction of the barrier term leads also to new features at the basic level. For instance, with the standard Morse potential (without including the barrier) a domain wall solution of the form Y (x) = ln[1 + e√2/S(x−x0)] can be obtained [21], where Y represents the separation of bases at position x and S set the wall width (and it is dependent on model parameters).This solution describes a configuration where one part of the molecule (x < x0) is closed and the other one (x ≫ x0) is opened and the bp separation grows linearly with space. However, this solution, for the standard PBD model, is unstable [21]. We show here that with the barrier term a stable domain wall solution can be found and thus, different oscillation modes corresponding to the different opening states of the chain can be obtained. In this paper, we will focus on the analysis of the unzipping process. We show that the inclusion of the barrier on the Morse potential allows us to obtain a stable domain wall solution for different values of the unzipping force. The stability analysis of this solution allows us to establish the frequency interval where resonant mechanisms can be found and also to estimate the escape rates for the unzipping. With these elements we finally study the combination of thermal noise and an external ac field in the unzipping process. The paper is developed as follows: model and methods are shown in section II. The linear stability analysis of the domain wall solution is presented in section III. The estimation of the escape rates is made in section IV and the influence of the ac field in the unzipping process is 3 presented in section V. II. MODEL AND METHODS We use the PBD model with the inclusion of a Gaussian barrier as in [7, 20]. Similar models with other expressions for the barrier were proposed in [18, 19]. The 1d Hamiltonian, in terms of the separation of bp yn, is given by the following expression: H = Xn (cid:20) p2 n 2m + V (yn) + W (yn, yn−1)(cid:21) , (1) where the first term corresponds to the kinetic energy (pn = mdyn/dt is the momentum of the nth bp and m its reduced mass) and the potential energy is given by the two other contributions: the on-site potential V (yn), that describes the interaction of bases belonging to the same pair; and the stacking interaction W (yn, yn−1), that gives the interaction of consecutive bps. V (yn) is defined as the sum of the Morse potential and a Gaussian barrier: V (yn) = D(e−αyn − 1)2 + Ge−(yn−y0)2/b. (2) D is the bp dissociation energy and α sets the width of the potential well. The Gaussian barrier models the entropic barrier that bps have to overcome to open and close again. The origin of this term comes from the fact that when the hydrogen bonds are broken, the bases can form new bonds with the solvent molecules and thus there must be an energetic cost to close them [18, 20]. The parameters G, y0 and b are the barrier height, position and width respectively. The term W (yn, yn−1) accounts for the stacking interactions: W (yn, yn−1) = 1 2 K(1 + ρe−δ(yn+yn−1))(yn − yn−1)2, (3) K is the coupling constant, ρ sets the anharmonic character of the stacking interaction and δ sets the scale length for this behavior. To simulate the unzipping under constant force, we added a term Vf = −F y1 to the first bp of the chain. A schematic view of the model and the potential V (yn) is depicted in figure 1. The values of the potential parameters are the same of those used in [7, 20] for a ho- mogeneous AT chain: D = 0.05185 eV and α = 4 A−1, for the Morse potential; G = 3D, y0 = 2/α and b = 1/2α2 for the Gaussian barrier; and K = 0.03 eVA−2, δ = 0.8 A−1 and 4 a) F W(y ,y n ) n−1 y n−1 ) n y ( V y n V(y ) n 7 6 5 4 3 2 1 0 b) 0 2 y n 4 6 8 FIG. 1: (Color online) a) Scheme of the unzipping process. b) Morse potential with the Gaussian barrier. ρ = 3, for the stacking interaction. For simplicity, we use dimensionless units: length is given in units of α−1; mass in units of the nucleotide mass (300 uma) and energy in units of D. With these transformations we get Fu = 332 pN, ωu = 5.15 rad/ps (0.82 THz) (and Tu = 602 K, for the force, frequency and temperature units, respectively. Given the shape of the potential V (yn), a stable domain wall solution can be obtained for different values of F . Beyond a critical value Fd, called the depinning force, the domain wall displaces from its equilibrium position and no static solution can be found. We carry out the linear stability analysis of this solution. For this, the configuration {yeq n } satisfying the equilibrium conditions has to fulfill: ∂(V (yn, F ) + W (yn, yn−1) + W (yn+1, yn)) ∂yn = 0. (4) where V (yn, F ) = V (yn) − F y1δ1n. To solve the nonlinear equation system 4, it is necessary to propose an initial guess for the solution. As we are interested in a domain wall solution, we start from a configuration with half of the chain at yn < y0 (closed state) and the other one at yn > y0 (open state). Using this initial guess, the nonlinear equation system 4 is solved. Once a solution is reached, we proceed to analyze its linear stability i.e. its behavior under small perturbation. To do that we build the Jacobian matrix of forces on each particle or, equivalently, the Hessian matrix of the potential energy [24]. The eigenvalues ω2 i obtained by diagonalizing the Hessian matrix of the system provide the signatures of the equilibrium 5 solution. If the solution is stable, all eigenvalues have to be positive. An unstable equilibrium solution can also be obtained for equation system 4 by using the following procedure: first, the stable equilibrium configuration yeq n is determined. Then, we add a small displacement to the bp on the close state that is closer to the open state (yeq i → yeq by yn = yeq i + dy). The system of equations 4 is solved again with the initial ansatz given n , n 6= i and yeq i as a parameter, i.e., the system has N − 1 equations. The energy of the obtained configuration E1 is computed. The same procedure is repeated by displacing yeq i i → yeq further, i.e., yeq i + sdy (s = 1, 2, ...), solving the system equations for the remaining y and calculating the energy Es for the obtained configuration. The procedure stops when a maximum value of the energy is reached. The configuration at this energy value corresponds to the unstable solution of the system. The eigenvalues of the stability matrix are all positive except one which corresponds to the unstable direction i.e. a saddle point in the energy landscape. If thermal noise or an external ac field are included, the Langevin equations of the motion are integrated numerically: m ∂2yn ∂t2 + mγ ∂yn ∂t = − ∂ [W (yn, yn−1) + W (yn+1, yn)] ∂yn ∂V ∂yn − + ξn(t) + Acos(ωt), (5) where m is the reduced mass of the bp, γ is the effective damping of the system, T is the temperature, A and ω are the amplitude and frequency field respectively. ξ(t) is a white Gaussian noise with hξn(t)i = 0 and hξn(t)ξk(t′)i = 2mγkBT δnkδ(t − t′). To integrate equation system 5 we use a stochastic Runge-Kutta algorithm [22, 23]. III. RESULTS A. Linear Stability Analysis of the domain wall solution We use a homogeneous AT chain of N = 10 bps. One end has fixed boundary conditions, i.e., yN +1 = 0 and the other one, where the force is applied, is free. We obtain the stable domain wall solution by solving 4 for different values of the constant force. The set of nonlinear equations is integrated numerically within Matlab with the trust-region-dogleg 6 ω 3 2 1 0 −1 −2 −3 0 0.5 1 1.5 2 F 2.5 3 3.5 4 ω 1 ω 2 ω 3 ω 4 ω 5 ω 6 ω 7 ω 8 ω 9 ω ω 10 u1 FIG. 2: (Color online) Frequency spectrum at different forces. ωu1 is the negative eigenvalue of the unstable solution. method. The initial guess for the solution is set with half of the chain at yi = 5, i = 1, .., 5 (open state) and the other one at yi = 0, i = 6, .., 10 (closed state). We start from F = 0 and then increase the force until no equilibrium configuration can be found. The force value at this point is defined as Fd. The unstable solution is also obtained by following the method depicted in the previous section. The eigenvalues ωi for the stable solution are shown in figure 2. The negative eigenvalue ωu1 for the unstable solution is also displayed. Two frequency bands are identified: one at 0 < ω < 0.5 and the other one at 1.5 < ω < 1.7, corresponding to the open and closed part of the chain, respectively. The position of these bands may be estimated analytically using the dispersion relations for the linear exci- tations inside the Morse potential (closed chain), mω2 ≈ 2Dα2 + 2K(1 + ρ)[1− cos(πn/N)]; and for a free Gaussian chain (open chain), mω2 = 2K[1− cos(πn/N)] [20]. Thus, the lower frequency band corresponds to the open chain while the upper one to the closed chain. One additional mode, corresponding to the domain wall between the open and the closed sides, is also observed. With our model parameters, the domain wall contains only one bp (the 6th bp). The location of this mode depends on the value of the applied force. At F < 1.1 and F > 4.1 the mode is located between the lower and the upper frequency band while at 7 F=1.2 F=3.5 F=4 F=4.13 F=0 F=0.6 10 5 0 −5 −10 2 y ∂ / ) y ( V 2 ∂ −15 −0.5 0 0.5 1 y 1.5 2 2.5 FIG. 3: Second derivative of the Morse potential and 6thbp positions at different forces. 1.1 < F < 4.1 is over the upper band. This may be explained because the second derivative of the Morse potential changes with position and ω2 is directly proportional to it (see figure 3). When F → Fd, the intermediate mode goes to the upper limit of the lower band. At F = 0, frequency values corresponding to the open part of the chain are larger than those obtained for F > 0. This is due to the presence of a local minimum beyond the barrier at y ≈ 3.75. When the force is applied, these modes approach zero indicating that bps are displaced away from that minimum. We have obtained Fd ≈ 4.13, which corresponds to a force of 1371 pN. This force value is much larger than the mean force reported for DNA unzipping (≈ 15 pN) [1]. However, we cannot compare these numbers since Fd is defined at zero temperature and no thermal effects are included. The real experiments are carried out at temperatures where thermal fluctuations help the opening of a bp. Thus, thermal fluctuations can reduce the unzipping forces as we discuss below. The eigenvectors corresponding to the stable domain wall solution at F = 2 are shown in figure 4 . A localized eigenvector at the 6th bp appears while the others are spread over the remaining bps. The eigenvectors corresponding to eigenvalues from λ1 to λ5 are spread over the five sites located in the open side of the chain and have zero amplitude over the remaining sites. They describe the modes where the open bps oscillate. In a similar way, eigenvectors associated to λ7 to λ10 describe the oscillations of the closed side of the chain. 8 1 0.5 0 −0.5 e d u t i l p m A . i g E λ 1 λ 5 λ 4 λ 3 λ 2 1 0.5 0 −0.5 e d u t i l p m A . i g E 1 0.5 0 −0.5 e d u t i l p m A . i g E 1 0.5 0 −0.5 e d u t i l p m A . i g E 1 0.5 0 −0.5 e d u t i l p m A . i g E −1 0 10 5 Index −1 0 10 5 Index −1 0 10 5 Index −1 0 10 5 Index −1 0 10 5 Index 1 0.5 0 −0.5 e d u t i l p m A . i g E λ 6 λ 10 λ 9 λ 8 λ 7 1 0.5 0 −0.5 e d u t i l p m A . i g E 1 0.5 0 −0.5 e d u t i l p m A . i g E 1 0.5 0 −0.5 e d u t i l p m A . i g E 1 0.5 0 −0.5 e d u t i l p m A . i g E −1 0 10 5 Index −1 0 10 5 Index −1 0 10 5 Index −1 0 10 5 Index −1 0 10 5 Index FIG. 4: Eigenvectors corresponding to stable solution at F = 2. The behavior of the eigenvectors corresponding to λ6 and λ7 (of the stable solution) and λu1 (of the unstable solution) at different forces are shown in figure 5. The unstable solution eigenvector remains localized at 6th bp for all values of F . The eigenvector corresponding to λ6 becomes localized after F > 1.1. Near the depinning force both eigenvectors coincide. As the domain wall is very sharp, the positions of the frequency bands as well as the Fd are almost the same if other sequence lengths and opening positions are used. This method can be extended to a homogeneous CG chain. The results are similar but the oscillation modes are obtained at different frequency values and the depinning force is larger. B. Estimation of the escape rates We can use these previous results to make some predictions on the thermal behavior of the unzipping process. One should be aware that the behavior of the domain wall solution can be approached by that of a single particle. This is because of the strong localization of such solution around the fork site [27, 28]. Indeed, in the presence of thermal noise, the unzipping process can be regarded as the problem of the escape of a particle from a well [21]. This problem has been widely studied and some analytical expressions for the escape 9 e d u t i l p m A . i g E e d u t i l p m A . i g E 1 0.5 0 −0.5 −1 0 1 0.5 0 −0.5 −1 0 F=0 λ 6 λ 7 λ u1 5 Index F=3.5 1 F=0.6 1 F=1.2 e d u t i l p m A . i g E 0.5 0 −0.5 10 −1 0 1 e d u t i l p m A . i g E 0.5 0 −0.5 5 Index F=4 e d u t i l p m A . i g E 0.5 0 −0.5 10 −1 0 1 e d u t i l p m A . i g E 0.5 0 −0.5 5 Index F=4.13 10 5 Index 10 −1 0 10 −1 0 5 Index 5 Index 10 FIG. 5: (Color online) Eigenvectors corresponding to λ6 and λ7 of the stable solution and λu1 of the unstable solution at different forces. rate ke have been proposed at different damping regimes [25, 26, 29]. In the moderate to high damping regime the following equation can be used: ke = {[1 + ( γ 2ωb )2]1/2 − γ 2ωb} ωa 2π e−E/kBT (6) where ωa and ωb are related to the curvature at the well and the barrier respectively and E is the energy of the barrier. From this formula ke is given in units of frequency (0.82 THz). In our case, the displacement of the domain wall corresponds to the escape of the particle from the well. ωa, ωb and E are the effective parameters of the well and the barrier that take into account the action of the remaining bps and the applied force. We take for wa the eigenvalue of the stable solution corresponding to the 6th bp; for wb the eigenvalue of the unstable solution and for E the height of the Peierls-Nabarro barrier which is the energy that the domain wall has to overcome to displace to the next bp site (which corresponds to the motion of 6th bp). This term is given by E = Eunstable − Estable, where Eunstable is the energy of the unstable solution and Estable is that of the stable one for a given value of F , figure 6. As expected, the value of E falls to zero at F = Fd. Near Fd, the dependence of E with F follows the relation E ∝ (1 − F/Fd)1.5 as shown in the inset of 6. This is the expected behavior for a single particle inside a well [30]. That shows that the approximation 10 of replacing the domain wall by a single particle should be good enough. The behavior of ωa and ωb as function of F have been shown in figure 2. 4 3.5 3 2.5 E 2 1.5 1 0.5 0 0 102 100 10−2 10−4 ) E ( g o l 10−6 10−3 10−2 10−1 log(1−F/Fd) 100 0.5 1 1.5 2 F 2.5 3 3.5 4 4.5 FIG. 6: Peierls-Nabarro energy at different forces. Inset: Fit of the E vs 1 − F/Fd near Fd. We estimate ke at different values of the applied force, damping and temperature (see figure 7). At low temperature, T = 0.1, ke is almost zero for forces below F = 1. Beyond this force value, there is an increase on ke because of the decrease on the barrier energy. Close to Fd, the escape rate falls abruptly although energy barrier tends to zero. This is due to the abrupt decreasing of ωa and ωb at this point. However, in these ranges of the force, expression 6 is unphysical. One expects to have reliable results for barriers E/kBT > 3 [31]. When temperature is increased, T = 0.5 (room temperature), the behavior of ke with the force is similar. As expected, the values are larger than those for T = 0.1. We look at the values of ke for F = 0.2 (which is of the order of experimental forces) for both temperatures and γ = 0.1. The obtained values are ke ∼ 10−16 and ke ∼ 10−4 for T = 0.1 and T = 0.5, respectively. This is the reason for which the value of Fd, defined at zero temperature, is high when compared with experiments. Indeed, DNA mechanical unzipping is an assisted thermal process dependent on the applied force. The values obtained for ke also suggest that other frequency dependent phenomena as the stochastic resonance or resonant activation would take place at frequencies lower than those obtained from the linear stability analysis. 11 a) 10−5 e k 10−10 10−15 0 b) 10−1 e k 10−2 10−3 10−4 0 γ=0.1 γ=1 γ=10 2 F 4 6 γ=0.1 γ=1 γ=10 2 F 4 6 FIG. 7: (Color online) Escape rates at different damping and temperature values. Units of ke are those of the frequency (0.82 THz). a) T = 0.1 b) T = 0.5 . C. Influence of the ac field and temperature in the unzipping Finally, we study the influence of an ac external field and thermal noise on the unzipping process. The frequencies values of the field used here are in the interval of the modes obtained in the previous section. These values are in the order of 1T Hz. The ac field is applied over all bps of the chain. Two damping γ = 0.1 and γ = 1, and four temperature values are used. The field amplitude is A = 0.4. This value is small when compared with the depinning force. The parameter values are in the same order of those used in [7] when studying the influence of an ac field in the melting transition and bubble formation on DNA. In these simulations we use a homogeneous AT chain of 20 bps and start from a closed state of the chain. The equations of motion are thermalized during a time ts = 1000 without force. After this time, both the constant force F and the ac field A are included in the motion equations and the system is integrated during ts = 600 (which corresponds approximately to 100 oscillation cycles of the external field). Then, we compute the temporal average of the domain wall position at different values of the frequency and the constant force F . F is varied from F = 0.1 until F = Fd. The results are averaged over 10 realizations. Figure 8 shows the mean domain wall position (on color scale) at different temperatures: T = 0, T = 0.1(60K), T = 0.4(240K), T = 0.5(300K). For γ = 0.1 and T = 0, an increase on domain wall position is obtained for frequency 12 T=0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 ω T=0.4 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 ω T=0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 ω T=0.4 4 3.5 3 2.5 2 1.5 1 0.5 F 4 3.5 3 2.5 2 1.5 1 0.5 F 4 3.5 3 2.5 2 1.5 1 0.5 F 4 3.5 3 2.5 2 1.5 1 0.5 F 20 18 16 14 12 10 8 6 4 2 20 18 16 14 12 10 8 6 4 2 20 18 16 14 12 10 8 6 4 2 20 18 16 14 12 10 8 6 4 2 4 3.5 3 2.5 2 1.5 1 0.5 F 4 3.5 3 2.5 2 1.5 1 0.5 F 4 3.5 3 2.5 2 1.5 1 0.5 F 4 3.5 3 2.5 2 1.5 1 0.5 F T=0.1 a) 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 ω T=0.5 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 ω T=0.1 b) 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 ω T=0.5 20 18 16 14 12 10 8 6 4 2 20 18 16 14 12 10 8 6 4 2 20 18 16 14 12 10 8 6 4 2 20 18 16 14 12 10 8 6 4 2 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 ω 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 ω FIG. 8: (Color online) Influence of ac field on the domain wall position at different temperatures (see text). a) γ = 0.1. b) γ = 1. Note that now the simulated chain consists of 20 sites. . values in the range 1.4 < ω < 1.8 (a maximum of the mean domain wall position is observed around these frequency values). These values correspond to those of the upper frequency band obtained previously, i.e., to the oscillation modes of bps inside the well of the Morse potential. The peak is slightly asymmetric which may be explained for the presence of the oscillation mode of the 6th bp. Thus, the mechanism of opening is resonant at these frequen- cies. As temperature increases the band spreads and the force for opening a given number of bp decrease because thermal noise assists opening events. This has been interpreted as a reduction of the critical force for unzipping due to resonance and temperature effects. For a larger damping γ = 1, there is little influence of the frequency on the forces to unzipping the chain. This is the behavior expected for a resonant mechanism. These results are in 13 agreement with [7], where maximum displacements were found around ω = 9rad/ps for a homogeneous AT chain and decreased when a higher damping was used. The oscillations modes corresponding to the open part of the chain do not influence the unzipping process. Simulations were also carried out at different field amplitudes (results not shown). In agree- ment with [7], the resonant influence of the field is observed for field amplitudes larger than a given threshold. F 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 F 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 T=0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 ω T=0.4 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 ω 20 18 16 14 12 10 8 6 4 2 20 18 16 14 12 10 8 6 4 2 F 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 F 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 T=0.1 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 ω T=0.5 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 ω 20 18 16 14 12 10 8 6 4 2 20 18 16 14 12 10 8 6 4 2 FIG. 9: (Color online) Influence of ac field on the domain wall position at different temperatures in the non barrier case and γ = 0.1. Note that now the simulated chain consists of 20 sites. Note also that force scale is different to than that used in 8. It is important to point out that the resonant influence of the ac field in the unzipping process, as well as in the melting and the bubble formation in the framework of the PBD model, is only possible with the inclusion of the barrier term in the standard Morse potential. When this term is absent, there is no resonant peak in the dependence on the frequency of the external ac field, at least in the interval we have studied. This is mainly due to the stability that the barrier confer to the domain wall solution. As we have seen, a frequency can be associated to the domain, that one corresponding to the localized eigenvector at the fork position. This frequency is in the range of the upper band of the linear stability spectrum. So, the domain wall frequency has no sense without the barrier and we do not expect to have resonant behavior. In order to check this prediction, we have performed a numerical simulation of the model without the barrier in the same range of parameters. Figure 9 shows the result for the same parameter values as in figure 8. Also, due to the lack of a barrier, the depinning force is reduced. Anyway, one can observe a drastic change in the 14 behavior respect to the frequency. No resonant reduction of the unzipping force is revealed. Thus, the inclusion of this term not only produces some results closer to experimental ones (sharper melting transition, wider and with larger lifetime bubbles) [18 -- 20] but also leads to new features that are absent in the model without the inclusion of the barrier. IV. CONCLUDING REMARKS In summary, we have studied the features of the unzipping process when a Gaussian barrier is added to the standard Morse potential and the influence of an external alternating field on the dynamics of this process. We have obtained the linear frequency modes corre- sponding to different opening states of the chain. The frequencies that influence resonantly in the unzipping process are those of the close part of the chain. We were also able to estimate the escape rates for the unzipping process under different parameters (temperature and damping). The knowledge of this parameter is important to understand force spec- troscopy experiments and other frequency dependent mechanisms as stochastic resonance and resonant activation. The analysis we have presented here may be extended to real chains if sequence dependent parameters are included. The inclusion of the barrier in the Morse potential leads to new features that are absent in the original PBD model with regard to the response to an ac external field. It should be stressed that also the inclusion of external fields provokes a dramatic reduction of the critical unzipping force at resonant frequencies. This could give a way for testing the validity of the inclusion of this term in a real experiment where this mechanism can be checked. Acknowledgments We thank J.J. Mazo for discussion and critical reading of the manuscript. This work is supported by the Spanish DGICYT Projects No. FIS2011-25167, co-financed by FEDER funds, and by the Comunidad de Arag´on through a grant to the FENOL group. AEBP 15 acknowledges the financial support of University of Zaragoza and Banco Santander. [1] B. Essevaz-Roulet, U. Bockelmann, and F. Heslot, Proc. Natl. Acad. Sci. USA 94 (1997) 11935. [2] U. Bockelmann, Ph. Thomen, B. Essevaz-Roulet, V. Viasnoff, and F. Heslot, Biophysical Journal 82 (2002) 1537. [3] C. Danilowicz, V. W. Coljee, C. Bouzigues, D. K. Lubensky, D. R. Nelson, and M. Prentiss, Proc. Natl. Acad. Sci. USA 100 (2003) 1694. [4] J. M. Huguet, C. V. Bizarro, N. Forns., S. B. Smith, C. Bustamante., and F. Ritort., Proc. Natl. Acad. Sci. USA 107 (2010) 15431. [5] B. S. Alexandrov, V. Gelev, A. R. Bishop, A. Usheva, and K. O. Rasmussen, Phys. Lett. A 374 (2010) 1214. [6] P. Maniadis, B. S. Alexandrov, A. R. Bishop, and K. O. Rasmussen, Phys. Rev. E 83 (2011) 011904. [7] A. E. Bergues-Pupo, J. M. Bergues, and F. Falo, Phys. Rev. E 87 (2013) 022703. [8] J. Bock, Y. Fukuyo, S. Kang, M. L. Phipps, L. B. Alexandrov, K. O. Rasmussen, A. R. Bishop, E. D. Rosen, J. S. Martinez, H. T. Chen, G. Rodriguez, B. S. Alexandrov, and A. Usheva, PLoS ONE 5 (2010) 15806. [9] O. Braun, A. Hanke, and U. Seifert, Phys.Rev.Lett. 93 (2004) 158105-1. [10] K. Hayashi, S. de Lorenzo, M. Manosas, J. M. Huguet and F. Ritort, Phys. Rev. X 2 (2012) 031012. [11] G. Mishra, P. Sadhukhan, S. M. Bhattacharjee, and S. Kumar, Phys. Rev. E 87, (2013) 022718. [12] T. Dauxois, M. Peyrard, and A. R. Bishop, Phys. Rev. E 47 (1993) 684. [13] S. Cuesta-Lopez, J. Errami, F. Falo, and M. Peyrard, J. of Biol. Phys. 31 (2005) 273. [14] N. K. Voulgarakis, A. Redondo, A. R. Bishop, and K. O. Rasmussen, Nano Letters 6 (2006) 1483. [15] N. K. Voulgarakis, A. Redondo, A. R. Bishop, and K. O. Rasmussen, Phys. Rev. Lett. 96 (2006) 248101. [16] A. Singh, B. Mittal, and N. Singh, Physics Express, 3 (2013) 18. [17] G. Weber, J. W. Essex, and C. Neylon, Nat. Phys. 5 (2009) 769. 16 [18] G. Weber, Europhys. Lett. 73 (2006) 806. [19] M. Peyrard, S. Cuesta-Lopez, G. James, J. Biol. Phys. 35 (2009) 73. [20] R. Tapia-Rojo, J. J. Mazo, and F. Falo, Phys. Rev. E 82 (2010) 031916. [21] M. Peyrard, Nonlinearity 17 (2004) R1. [22] H. S. Greenside and E. Helfand, The Bell System Technical Journal 60 (1981) 1927. [23] E. Helfand, Bell System Technical Journal 58 (1979) 2289. [24] S. H. Strogatz, N on Linear Dynamics an Chaos, Perseus Books (1994) Reading, Mas- sachusetts. [25] H. Kramers, Physica 7 (1940) 284. [26] P. Hanggi, P. Talkner, and M. Borkovec, Rev. Mod. Phys. 62 (1990) 251. [27] P.J. Martinez, F. Falo, J.J. Mazo, L.M. Floria and A. Sanchez Phys. Rev. B 56 (1997) 87. [28] J. J. Mazo, F. Naranjo, and K. Segall, Phys. Rev. B, 78 (2008) 174510. [29] J.J. Mazo, O.Y. Fajardo, and D. Zueco, J. Chem. Phys. 138 (2013) 104105. [30] T. A. Fulton, and L. N. Dunkleberger, Phys. Rev. B 9 (1974) 4760. [31] J.J. Mazo, F. Naranjo, and D. Zueco, Phys. Rev. B 82 (2010) 094505. 17
1307.3563
1
1307
2013-07-12T20:04:15
Locomotion of helical bodies in viscoelastic fluids: enhanced swimming at large helical amplitudes
[ "physics.bio-ph", "cond-mat.soft", "physics.flu-dyn" ]
The motion of a rotating helical body in a viscoelastic fluid is considered. In the case of force-free swimming, the introduction of viscoelasticity can either enhance or retard the swimming speed and locomotive efficiency, depending on the body geometry, fluid properties, and the body rotation rate. Numerical solutions of the Oldroyd-B equations show how previous theoretical predictions break down with increasing helical radius or with decreasing filament thickness. Helices of large pitch angle show an increase in swimming speed to a local maximum at a Deborah number of order unity. The numerical results show how the small-amplitude theoretical calculations connect smoothly to the large-amplitude experimental measurements.
physics.bio-ph
physics
Locomotion of helical bodies in viscoelastic fluids: enhanced swimming at large helical amplitudes Department of Mathematics, University of Wisconsin, 480 Lincoln Drive, Madison, WI 53706, USA Saverio E. Spagnolie∗ School of Engineering, Brown University, 182 Hope Street, Providence, RI 02912, USA Bin Liu Department of Physics and School of Engineering, Brown University, 182 Hope Street, Providence, RI 02912, USA Thomas R. Powers† (Dated: April 5, 2018) 3 1 0 2 l u J 2 1 ] h p - o i b . s c i s y h p [ 1 v 3 6 5 3 . 7 0 3 1 : v i X r a The motion of a rotating helical body in a viscoelastic fluid is considered. In the case of force-free swim- ming, the introduction of viscoelasticity can either enhance or retard the swimming speed and locomotive effi- ciency, depending on the body geometry, fluid properties, and the body rotation rate. Numerical solutions of the Oldroyd-B equations show how previous theoretical predictions break down with increasing helical radius or with decreasing filament thickness. Helices of large pitch angle show an increase in swimming speed to a local maximum at a Deborah number of order unity. The numerical results show how the small-amplitude theoretical calculations connect smoothly to the large-amplitude experimental measurements. Much has been learned about the swimming of microorgan- isms in viscous environments over the last decade [1]. The pe- culiar behavior of complex fluids has also seen a recent burst of renewed interest, particularly as applied to biological sys- tems [2]. Progress in both fields has begun to blur together, since many organisms commonly swim in shear-thinning or viscoelastic fluids. Some of those fluids are complex specif- ically because of suspensions of microorganisms swimming and diffusing throughout [3–7]. Examples of microorgan- isms swimming in complex fluid environments include mam- malian spermatozoa through cervical fluid [8], the Lyme dis- ease spirochete B. burgdorferi through the extracellular ma- trix of our skin [9, 10], and the nematode C. elegans in water- saturated soil [11]. Organisms such as H. pylori have even been found to reduce fluid elasticity in order to swim through mucus [12]. A puzzle has recently emerged in the study of swimming through complex fluids. Theory, experiment, and simula- tion have indicated the possibility of both enhancement and retardation of swimming speeds in viscoelastic fluids (see Fig. 1a-f). Helically-shaped bacteria such as Leptospira and B. burgdorferi swim faster in solutions with methylcellu- lose than in non-viscoelastic solutions of the same viscos- ity [9, 13]. C. elegans, however, which propels itself by pla- nar body undulations, swims slower in a viscoelastic fluid than in a viscous fluid [14]. Spermatozoan cells swim slower when the fluid has an elastic response, but along straighter paths due to resultant changes in the flagellar shape, with hy- peractivated spermatozoan cells swimming faster than normal cells [15]. The consequences of fluid viscoelasticity on swim- ming is not, then, a question that can be answered broadly; effects appear to depend sensitively upon the geometry of the swimming stroke and the rheology of the complex fluid. There have also been a number of recent analytical, nu- merical, and scale-model explorations. Analysis is commonly performed on the Oldroyd-B equations, which describe a vis- coelastic flow with no shear-thinning or thickening [16, 17]. Using the Oldroyd-B model and others, Lauga showed that an infinite sheet passing small amplitude waves always swims slower with the introduction of viscoelasticity [18]. An iden- tical factor of swimming speed reduction was recovered by Fu et al. for a nearly cylindrical body of small pitch angle when passing helical waves [19], and similarly for the passage of planar waves [20]. Teran et al. showed that finite undulatory bodies of large wave amplitude can swim faster in a viscoelas- tic fluid [21], while Curtis and Gaffney showed the same for a three-sphere swimmer [22], as did Espinosa-Garcia et al. for flexible swimmers [23]. Finally, Liu et al. studied experimen- tally the motion of a rotating, force-free helical filament in a (viscoelastic) Boger fluid, finding that the swimming speed increased or decreased with viscoelasticity depending on the body geometry and rotation rate (see Fig. 1f) [24, 25]. In this paper we bridge the gap between the analytical pre- dictions and the experimental and numerical observations just described. By studying numerically the swimming of a helical body in an Oldroyd-B model fluid, we show that the theoreti- cal efforts do indeed capture the effects of viscoelasticity when the helical pitch angle is small and the filament radius is large: namely, that the swimming speed of a rotating helical body in this regime is always smaller than the same in a Newtonian fluid. We will then show how these theories break down for helices of large pitch angle and that the swimming speed can increase with the introduction of viscoelasticity. The results may improve our understanding of mammalian fertility and the spread of bacterial infections and diseases [10]. The paper is organized as follows. We begin by presenting the mathematical model of a rotating, force-free helical body in an Oldroyd-B fluid, followed by a dimensional analysis and a description of the numerical method. Helices of small pitch angle are explored, for which we show a continuous depar- ture of the small amplitude analytical theories from the results of the full model. We then turn to helices of large pitch an- gle where fluid elasticity is shown to increase the swimming speed to a local maximum at a Deborah number of order unity. 2 FIG. 1. (Color online) Experiments, theories, and simulations of swimming in viscoelastic fluids. With increasing fluid elasticity, (a) the helical bacterium Leptospira swims faster (see also Fig. 7 of [26]) [9, 13, 27]; (b) the nematode C. elegans swims slower [14, 28]; (c) a two-dimensional swimming sheet of small wave amplitude swims slower [18]; (d) a nearly-cylindrical swimmer passing helical waves of small pitch angle swims at the same speed as a two-dimensional sheet [19]; (e) a finite undulatory swimmer swims faster (instantaneous body velocity and mean-squared polymer distention field are shown [21, 29]); (f) and a thin helical body of arbitrary length can swim faster or slower, depending on the geometry and rotation rate [24]. (g) The axial component of fluid velocity generated by a rotating, force-free helical filament is shown; a helical fluid volume external to the coil is carried upward with the translating body, while a helical fluid volume internal to the coil is shuttled downward. Both helical waves and rigid body motions are considered, and the locomotive efficiency is addressed. We conclude with a summary and a discussion of future directions. The experiments of Liu et al. suggest that the force-free swimming speed of a helical filament is broadly independent of its length [24]. We therefore consider a right-handed helical body of infinite length with centerline x(s, t) = b[cos(ks + Ωt)x + sin(ks + Ωt)y] + ([1− (bk)2]1/2s + U∗t)z. Here, b is the helical radius, k is the wavenumber, s is the arc-length, Ω is a fixed rotation rate, and U∗ is the swimming speed. The body is shaped such that the boundary in a cross-sectional plane perpendicular to the z axis is circular with radius A/k. The distinction between helical waves and rigid body rotation is a rotation of this circle about the centerline in the latter. At the length and velocity scales relevant for microorgan- isms, viscous effects dominate inertial effects [1] and the mo- mentum and mass conservation equations are ∇p = ∇ · τ and ∇ · u = 0, where p is the pressure and τ is the total devi- atoric stress tensor. In the Oldroyd-B model, τ is the sum of a Newtonian solvent contribution, τs = ηs γ, and an ex- tra polymeric contribution, τp, where ηs is the solvent vis- cosity, γ = ∇u + ∇uT is the rate-of-strain tensor, and u is the fluid velocity. Meanwhile, τp is described by an upper- convected Maxwell model in which a single elastic relaxation timescale, λ1, and a viscous retardation timescale, λ2, are in- troduced (with λ2 = λ1ηs/η < λ1 and η the total zero shear rate viscosity) [16, 17]. Scaling lengths on 1/k, time on 1/Ω, ve- locities on Ω/k, and stresses on ηΩ, the total deviatoric stress is found to satisfy the dimensionless constitutive relation: τ + De (cid:79) τ = γ + (ηs/η)De (cid:79) γ. (1) Here we have defined the dimensionless Deborah number, De = λ1Ω, which compares the body rotation rate to the rate of elastic relaxation. The upper convective time derivative in τ = τt + u · ∇τ − ∇uT · τ − τ · ∇u. More (1) is defined as complex models include features such as multiple relaxation timescales and finite polymer extensibility [16]. (cid:79) The Deborah numbers relevant to microorganism motility are likely to span a very wide range. Relaxation times of cer- vical mucus have been measured from λ1 = 0.03 s to λ1 = 100 s; spermatozoa in this environment undulate with frequencies between 20 and 50 Hz, corresponding to Deborah numbers that are O(1) or much larger (see [18, 30]). An intriguing ex- ample of Deborah number variation is exhibited in the swim- ming of H. pylori, which rotates a helical flagellum at up to 5 Hz in viscoelastic mucus with a relaxation time of approxi- mately 100 s, but releases pH-increasing urease enzymes that decreases the local relaxation time to nearly 0.05 s, thereby decreasing the Deborah number from O(1000) to O(1) [12]. Due to the interaction with the fluid, the helical filament translates along the axial direction with dimensionless speed U, as illustrated in Fig. 2a. A no-slip condition is assumed on the body surface, and for computational purposes we place the filament inside a very large container where we set u = 0. The container is made sufficiently large so that further increases in its size have a negligible effect on the reported results. The problem is closed by requiring that the axial component of force on the body is zero. The constant swimming speed U is determined by assuming that a locomotive steady state has been achieved and by exploiting helical symmetry: the flow and stress fields everywhere are given by translation and rota- tion of the flow field through the z = 0 plane. Conversion to a helical coordinate system allows for z derivatives to be writ- ten as planar derivatives on z = 0. In a periodic steady state, time derivatives may be written as z derivatives, and hence (f)⌦(a)(b)(d)(g)(c)(e) 3 FIG. 2. (Color online) (a) Schematic of the dimensionless setup. The body rotates counter-clockwise with unit angular speed when seen from above and swims in the axial direction with dimensionless speed U. (b) Helical-wave swimming speed (normalized by the Newtonian swimming speed) of filaments of varying thickness, A = 2n−2. Here ψ = π/40 and ηs/η = 0.5. Deviations from the theoretical result of Fu et al. [19] (dotted line) are logarithmic in A. (c) Viscoelasticity leads to faster swimming for helices of sufficiently large pitch angle. The filament thickness is fixed, A = 0.5. Solid lines denote helical waves, dashed lines denote rigid body rotation, and symbols denote pitch angle, ψ = π/40, π/10, and π/5. (d) Larger swimming speeds are achieved by thinner filaments of the same pitch angles as in (c); here A = 0.2. (e) The normalized swimming efficiency for helical waves and rigid body rotations. Symbols denote the same helices as in (c). by planar derivatives. The Oldroyd-B equations are solved numerically using a mixed pseudo-spectral / finite differences approach. The mathematical model and numerical method are described in greater detail in the supplementary material. We first compare the numerical results to the analytical work of Fu et al. [19] in the case of a helical wave with small pitch angle. For ψ = π/40, the normalized swimming speed is shown in Fig. 2b for a range of Deborah numbers and fila- ment sizes. Here as in the remainder of the paper we fix the viscosity ratio to ηs/η = 0.5. Each solid line corresponds to a different filament radius, A = 2n−2, for n = 0, 1, ..., 6. By in- creasing the filament thickness the swimming speed converges monotonically to the analytical result, shown as a dashed line. Viscoelasticity in this case decreases the swimming speed of helices with small pitch angles, even for slender bodies, con- trasting with the enhanced speeds predicted in a viscous fluid in the presence of stationary obstacles [31]. The departure of the results from the analytical theory are logarithmic in A, consistent with the analytical development [19]. The analytical results at small pitch angle have been recov- ered, but can the increased swimming speeds seen in experi- ments be found? Figure 2c shows the normalized swimming speed as a function of the Deborah number for three different pitch angles, ψ = π/40, π/10, and π/5, with A = 0.5. Heli- cal wave and rigid body rotation results are shown as solid and dashed lines, respectively. For small Deborah numbers we ob- serve U/UN = 1 + O(De2), as required by symmetry. In both cases, for a given Deborah number in the regime considered, the swimming speed increases as the pitch angle is increased to ψ = π/5. Moreover, beyond a critical pitch angle we find a range of Deborah numbers for which the swimming speed is larger in a viscoelastic fluid than in a Newtonian fluid, just as observed in experiments [24]. Rigid body motion, which generates an extra rotational flow around the helical filament as compared to helical waves, am- plifies the effects of viscoelasticity, particularly for small pitch angles where rotational flow is dominant. Viscoelastic effects are amplified with decreasing ηs/η as well (not shown). Fil- ament thickness is also important; Fig. 2d shows the helical- wave swimming speeds of slender filaments (A = 0.2), which are greater than those shown in Fig. 2c. In particular, for the intermediate pitch angle ψ = π/10, reducing the filament thickness introduces a regime in small Deborah number for which the relative swimming speed is greater than unity. (cid:82) A microorganism may benefit by swimming with greater efficiency rather than greater speed. We evaluate a com- mon measure of swimming efficiency, E = U2/P, where z=0 γ : τ dS is the rate of energy dissipation in P = (1/2) the fluid per unit length [18, 19, 21]. The results are shown in Fig. 2e for the same helical shapes considered in Fig. 2c. U/UN0.90.80.71.051.50.5(c)DeU/UNψ=π/40(b)1.50.5De\x{FFFF}1+(ηs/η)De21+De2\x{FFFF}A=2n−20.70.6n=01234560.9HelicalwaveRigidbodyrotationz=0(a)Uψ2πcos(ψ)xyz2A(d)U/UN0.51.5De0.91.05E/EN0.51.5De0.60.80.41.21.41.6(e)1.01.01.01.00.00.01.01.01.02.00.850.00.00.950.81.0 An important distinction between the passage of helical waves and rigid body rotation is observed. For rigid body rotation, the work done on the fluid does not vary dramatically with Deborah number; just as was found for planar undulations, the swimming speed is a proxy for swimming efficiency in the case of rigid body rotation [21]. The rotation of helices of large pitch angle therefore presents a means of swimming that is both faster and more efficient in a viscoelastic fluid when the rotation rate is properly tuned to the fluid environ- ment. For helical waves, however, the swimming speed is not so clearly linked to the efficiency; for ψ = π/10 the rela- tive swimming speed decreases with De while the relative ef- ficiency increases. By inspection of the no-slip boundary con- dition, the differences between rigid body rotation and helical waves are expected to diminish rapidly with both decreasing filament size and increasing pitch angle. FIG. 3. (Color online) The mean-squared polymer distention fields, tr(τp), from the passage of helical waves are shown for three Deb- orah numbers, for filament thickness A = 0.2 and pitch angles (a) ψ = π/40 and (b) ψ = π/5. The swimming speed decreases with Deborah number in (a) and increases in (b). The path of the body during rotation is indicated by a white dashed line in the bottom left panel. The largest polymer distention appears along the sec- ond/fourth quadrants in the former case, and along the first/third quadrants in the latter case, with the regions of maximal distention indicated by symbols. The effects of viscoelasticity on the swimming of helices are not easily predicted by thought experiments. The flow field created by a rotating helix in a Newtonian fluid is intri- cate; the extra polymeric stresses that develop due to this flow field, the response of the flow field to the polymeric stresses, and the interaction of solvent and polymeric forces with the helix all provide for a complex and highly nonlinear system. We do, however, observe a distinction in the polymeric stresses between cases where viscoelasticity either reduces or increases the swimming speed. Consider the spatial distribu- tion of tr(τp) for the passage of helical waves along two dif- ferent helical geometries. This quantity measures the mean- squared distention of the elastic polymers, and is shown for helices of small and large pitch angle in Fig. 3. For small Deborah numbers, the polymers are stretched most in regions of large fluid shear, which are at the leftmost and rightmost 4 points of the circular boundaries in Fig. 3 (the inner and outer edges of the filament, respectively). Fluid shear is largest in these regions due not only to the motion of the filament through the z = 0 plane, but also to the arrangement of the axial fluid velocity, as shown for De = 0 in Fig. 1g. Increasing the Deborah number, however, shows a distinct difference in the organization of polymeric stress in these two cases. For the helix of smaller pitch angle the primary re- gions of extra stress rotate clockwise. As these regions are displaced, they affect the underlying flow field, and both con- spire to reduce the filament swimming speed. For the helix of larger pitch angle, however, we observe the opposite shift: ex- tra polymeric stresses have shifted counter-clockwise. Much as in the case of flow past a cylinder, the reorganization of elastic fluid stress acts to shift the distribution of pressure, which further contributes to adjustments in both vertical and horizontal fluid forces on the body [32]. Recall that a Deborah number of order unity indicates a he- lical rotation rate comparable to the rate of elastic relaxation. When De ≈ 1, polymers distended by the motion of the body release stored elastic energy on a special timescale. Namely, a timescale such that the body can revisit the viscoelastic wake created on each pass through the same location. For helices of small pitch angle, the motion of the body in the z = 0 plane is muted, and fluid parcels do not make large excursions in the plane. With a larger pitch angle, the filament travels on a much wider circuit relative to its thickness, the body can in- teract with its own viscoelastic wake, and elastic stresses can be transmitted from the wake back onto the body upon its re- turn. This heuristic also suggests that a smaller filament thick- ness (with pitch angle fixed) may lead to increased swimming speeds, as we have already observed in Fig. 2d. We have shown that the introduction of viscoelasticity can either decrease or increase the force-free swimming speed and swimming efficiency of a rotating helical filament, depending on the helical geometry, the material properties of the fluid, and the body rotation rate. The results of our investigation connect the small amplitude theories, experimental observa- tions, and numerical investigations of the biological and me- chanical experiments shown in Fig. 1. Our findings may add context to the recent discovery that H. pylori reduces mu- cus elasticity to a range more suitable for effective locomo- tion [12]. Future work will explore the effects of viscoelastic- ity on the propulsion of elastic helical filaments, on flagellar bundling, and on more intricate solid structures such as those observed in bacterial polymorphism [33–35]. Similar behav- iors are expected in a nearby fluid pumping problem, where fluid is transported downward either faster or slower than the same in a Newtonian setting. We acknowledge helpful discussions with M. Graham, M. Shelley, R. Ewoldt, and G. Forest, and we are especially grate- ful to K. Breuer for ongoing collaboration. This work was supported by the NSF Grant No. CBET-0854108. De=1.50.0100.4 =⇡/40 =⇡/50tr(⌧p)tr(⌧p)U/UN>0U/UN<0De=0.3De=0.6xxxy(b)(a)001111011011011011011011011011011011xxxyyyyy ∗ [email protected] † thomas [email protected] [1] E. Lauga and T.R. Powers. The hydrodynamics of swimming microorganisms. Rep. Prog. Phys., 72:096601, 2009. [2] T. G. Squires, T. M. & Mason. Fluid mechanics of microrheol- ogy. Annu. Rev. Fluid Mech., 42:413–438, 2010. [3] J. O. Pedley, T. J. & Kessler. Hydrodynamic phenomena in suspensions of swimming microorganisms. Annu. Rev. Fluid Mech., 24:313–358, 1992. [4] T. J. Ishikawa, T. & Pedley. The rheology of a semi-dilute sus- pension of swimming model micro-organisms. J. Fluid Mech., 588:399–435, 2007. [5] S. Rafaı, L. Jibuti, and P. Peyla. Effective viscosity of mi- croswimmer suspensions. Phys. Rev. Lett., 104:098102, 2010. [6] D. Saintillan. The dilute rheology of swimming suspensions: A simple kinetic model. Exp. Mech., 50:1275–1281, 2010. [7] G. Foffano, J. S. Lintuvuori, A. N. Morozov, K. Stratford, M. E. Cates, and D. Marenduzzo. Bulk rheology and microrheology of active fluids. EPJE, 35:1–11, 2012. [8] S. S. Suarez and A. A. Pacey. Sperm transport in the female reproductive tract. Hum. Reprod. Update, 12:23–37, 2006. [9] R.B. Kimsey and Spielman A. Motility of the Lyme disease spirochetes in fluids as viscous as the extracellular matrix. J. Infect. Dis., 162:1205–1208, 1990. [10] M. W. Harman, S. M. Dunham-Ems, M. J. Caimano, A. A. Belperron, L. K. Bockenstedt, H. C. Fu, J. D. Radolf, and C. W. Wolgemuth. The heterogeneous motility of the Lyme disease spirochete in gelatin mimics dissemination through tis- sue. PNAS, 109:3059–3064, 2012. [11] S. Jung. Caenorhabditis elegans swimming in a saturated par- ticulate system. Phys. Fluids, 22:031903, 2010. [12] J.P. Celli, B.S. Turner, N.H. Afdhal, S. Keates, I. Ghiran, C.P. Kelly, R.H. Ewoldt, G.H. McKinley, P. So, S. Erramilli, et al. Helicobacter pylori moves through mucus by reducing mucin viscoelasticity. PNAS, 106:14321–14326, 2009. [13] L. Berg, H.C. & Turner. Movement of microorganisms in vis- cous environments. Nature, 278:349–351, 1979. [14] P. E. Shen, X. N. & Arratia. Undulatory swimming in viscoelas- tic fluids. Phys. Rev. Lett., 106:208101, 2011. [15] S. S. Suarez and X. Dai. Hyperactivation enhances mouse sperm capacity for penetrating viscoelastic media. Biol. Re- prod., 46:686–691, 1992. [16] R.B. Bird, R.C. Armstrong, and O. Hassager. Dynamics of poly- meric liquids. Vol. 1: Fluid mechanics. John Wiley and Sons Inc., New York, NY, 1987. 5 [17] R. G. Larson. The Structure and Rheology of Complex Fluids, volume 1. Oxford University Press, 1999. [18] E. Lauga. Propulsion in a viscoelastic fluid. Phys. Fluids, 19:083104, 2007. [19] H. C. Fu, C. W. Wolgemuth, and T. R. Powers. Swimming speeds of filaments in nonlinearly viscoelastic fluids. Phys. Flu- ids, 21:033102, 2009. [20] H. C. Fu, T. R. Powers, and H. C. Wolgemuth. Theory of swimming filaments in viscoelastic media. Phys. Rev. Lett., 99:258101–258105, 2007. [21] J. Teran, L. Fauci, and M. Shelley. Viscoelastic fluid response can increase the speed and efficiency of a free swimmer. Phys. Rev. Lett., 104:038101, 2010. [22] M. P. Curtis and E. A. Gaffney. Three-sphere swimmer in a nonlinear viscoelastic medium. Phys. Rev. E, 87:043006, 2013. [23] J. Espinosa-Garcia, E. Lauga, and R. Zenit. Fluid elasticity increases the locomotion of flexible swimmers. Phys. Fluids, 25:031701, 2013. [24] B. Liu, T. R. Powers, and K. S. Breuer. Force-free swim- ming of a model helical flagellum in viscoelastic fluids. PNAS, 108:19516–19520, 2011. [25] M. Dasgupta, B. Liu, H. C. Fu, M. Berhanu, K. S. Breuer, T. R. Powers, and A. Kudrolli. Speed of a swimming sheet in New- tonian and viscoelastic fluids. Phys. Rev. E, 87:013015, 2013. [26] A. R. Bharti, J. E. Nally, J. N. Ricaldi, M. A. Matthias, M. M. Diaz, M. A. Lovett, P. N. Levett, R. H. Gilman, M. R. Willig, E. Gotuzzo, et al. Leptospirosis: a zoonotic disease of global importance. Lancet Infect. Dis., 3:757–771, 2003. [27] Courtesy CDC, Janice Haney Carr, 2013. [28] Courtesy of X. Shen and P. Arratia, 2013. [29] Courtesy of J. Teran, L. Fauci, and M. Shelley, 2013. [30] D. J. Smith, E. A. Gaffney, H. Gadelha, N. Kapur, and J. C. Kirkman-Brown. Bend propagation in the flagella of migrating human sperm, and its modulation by viscosity. Cell Mot. and the Cyto., 66(4):220–236, 2009. [31] A. M. Leshansky. Enhanced low-Reynolds-number propul- sion in heterogeneous viscous environments. Phys. Rev. E, 80:051911, 2009. [32] N. Phan-Thien and H. S. Dou. Viscoelastic flow past a cylinder: drag coefficient. Comput. Meth. Appl. M., 180:243–266, 1999. [33] C. R. Calladine. Construction of bacterial flagella. Nature, 255:121–124, 1975. [34] S. V. Srigiriraju and T. R. Powers. Continuum model for poly- morphism of bacterial flagella. Phys. Rev. Lett., 94:248101, 2005. [35] S. E. Spagnolie and E. Lauga. Comparative hydrodynamics of bacterial polymorphism. Phys. Rev. Lett., 106:058103, 2011.
1006.5737
2
1006
2010-10-17T21:17:54
Stochastic effects in a seasonally forced epidemic model
[ "physics.bio-ph", "cond-mat.stat-mech", "nlin.AO", "q-bio.PE" ]
The interplay of seasonality, the system's nonlinearities and intrinsic stochasticity is studied for a seasonally forced susceptible-exposed-infective-recovered stochastic model. The model is explored in the parameter region that corresponds to childhood infectious diseases such as measles. The power spectrum of the stochastic fluctuations around the attractors of the deterministic system that describes the model in the thermodynamic limit is computed analytically and validated by stochastic simulations for large system sizes. Size effects are studied through additional simulations. Other effects such as switching between coexisting attractors induced by stochasticity often mentioned in the literature as playing an important role in the dynamics of childhood infectious diseases are also investigated. The main conclusion is that stochastic amplification, rather than these effects, is the key ingredient to understand the observed incidence patterns.
physics.bio-ph
physics
Stochastic effects in a seasonally forced epidemic model 1Centro de F´ısica da Mat´eria Condensada and Departamento de F´ısica, Faculdade de Ciencias da Universidade de Lisboa, P-1649-003 Lisboa Codex, Portugal 2Departamento de Engenharia Civil, Instituto Superior de Engenharia de Lisboa, P-1959-007 Lisboa Codex, Portugal G. Rozhnova1, 2 and A. Nunes1 The interplay of seasonality, the system's nonlinearities and intrinsic stochasticity is studied for a seasonally forced susceptible-exposed-infective-recovered stochastic model. The model is explored in the parameter region that corresponds to childhood infectious diseases such as measles. The power spectrum of the stochastic fluctuations around the attractors of the deterministic system that describes the model in the thermodynamic limit is computed analytically and validated by stochastic simulations for large system sizes. Size effects are studied through additional simulations. Other effects such as switching between coexisting attractors induced by stochasticity often mentioned in the literature as playing an important role in the dynamics of childhood infectious diseases are also investigated. The main conclusion is that stochastic amplification, rather than these effects, is the key ingredient to understand the observed incidence patterns. PACS numbers: 87.10.Mn, 87.10.Ca, 05.10.Gg, 05.45.-a I. INTRODUCTION Epidemic spread in human populations is a complex phenomenon whose comprehensive modeling has been a standing challenge for many years [1]. Several ingredients such as host population contact structure, host hetero- geneity, transmission mechanisms, interplay between im- mune response and pathogen evolution or demographic and environmental factors have been identified as play- ing an important role in the short and in the long term behavior of infection spread [2 -- 13]. Scarcity of data, es- pecially for long term behavior, and model parameters that are hard to measure accurately add to the complex- ity of the problem, making it difficult to identify the key ingredients of a parsimonious model for a given disease or class of diseases [14, 15]. Childhood infectious diseases have often been taken as a case study and model testing ground, because decades long of fairly well time resolved data records are avail- able, on one hand, and because of their different phe- nomenology despite the similarities in contagion mecha- nisms and in infectious, latent and immunity waning typ- ical times [16]. The common modeling approach for this class of diseases is a SEIR (susceptible-exposed-infective- recovered) compartmental model, with a periodic forcing that represents seasonal environmental or social factors that influence the transmission of the disease [16 -- 20]. Deterministic models based on this approach, where the role of stochasticity is merely that of making the system switch between coexisting attractors, successfully repro- duce the main features of the observed time series for measles incidence, but fail conspicuously to model the behavior of other childhood diseases that exhibit non- seasonal sustained oscillations on several long term data records [21, 22]. This failure has been addressed in the literature by claiming that this different dynamics may be either the result of more realistic latent and infectious period distributions or the evidence of stochastic effects that would show up as noisy oscillations with the same frequency as the damped oscillations of the determinis- tic system [15, 22]. The idea that stochastic effects may play a more fundamental role has also been discussed in the epidemiological literature [14, 23], prompting several recent analytical studies, all of which deal with unforced systems [24 -- 26]. Sustained oscillations typical of the incidence patterns of childhood infectious diseases are one of the features of the long term behavior of these unforced stochastic models. The power spectrum of the fluctuations around the endemic equilibrium computed analytically using van Kampen's system size expansion has well defined res- onant like peaks [24], which means that for moderate system sizes demographic stochasticity will generate sus- tained noisy oscillations, a phenomenon dubbed stochas- tic amplification when it was first studied in ecological and epidemiological models [24, 27]. Indeed, it has been shown that intrinsic noise enhanced by correlations in un- forced epidemic models may give rise to oscillations that are comparable to those due to seasonal forcing in de- terministic systems [28 -- 30]. It has also been shown that the dominant frequency of these stochastic fluctuations may differ significantly from the characteristic frequency of the deterministic system, especially in the presence of correlations [28, 29, 31]. Adding seasonality to these unforced systems may also give rise to significant non trivial effects in their long term behavior. Therefore, the analytical results for the fluctuations power spec- trum must be extended to the corresponding periodically forced models in order to assess the role of stochastic ef- fects in childhood diseases epidemiology through the in- terplay between seasonality, the system's nonlinearities and intrinsic stochasticity. The method developed in Ref. [27] was extended to fluctuations around cycles of forced or unforced systems in a series of recent papers [32, 33]. Here, we apply this method to a seasonally forced SEIR model in a realistic parameter region for childhood infectious diseases where different attractors exist or coexist. Since very little is known about the amplitude of seasonal forcing, we leave this as a free parameter in our study and consider sepa- rately the low, intermediate and high forcing regimes. In all cases, we find an excellent agreement between the an- alytical power spectra and the results of stochastic simu- lations. We use the latter to assess the role of competing attractors in explaining the observed time series of child- hood diseases epidemics, and we use the former to predict the number and position of the dominant non-seasonal peaks as a function of the epidemiological parameters. II. THE SEASONALLY FORCED DETERMINISTIC SEIR EPIDEMIC MODEL In this section, we will review a seasonally forced de- terministic SEIR model and its dynamics. Following most of the literature on childhood infectious diseases' modeling, we will take measles as a paradigmatic exam- ple throughout the paper, and the behavior of the sys- tem will be illustrated in the relevant parameter region [14, 17, 19]. Extensions of the SIR/SEIR model have also been considered, especially in the mathematical litera- ture, in connection with acute infectious diseases' mod- eling. These extensions may include, for instance, higher order nonlinearities in the infection term [34 -- 36], immi- gration of infectives [24, 37] or age structure [38]. How- ever, it is generally accepted that the seasonally forced SEIR model should capture the main features of the dy- namics of childhood infections. Consider then a homogeneously mixed population of constant size consisting of four classes of individuals: susceptibles (healthy individuals capable of contracting the infection), exposed (infected but not yet infectious individuals), infectives (infectious individuals capable of transmitting the infection) and recovered (permanently immune individuals). The dynamics of the SEIR model is governed by the following processes: 1) The susceptible individuals catch the infection from the infective individuals at a time dependent contact (or transmission) rate β(t). To incorporate seasonality in transmission of the infection, the contact rate is assumed to be a periodic sinusoidal function with period 1 year β(t) = β0(1 + β1 cos 2πt), (1) where t is measured in years. In Eq. (1), β1 ∈ [0, 1] is the amplitude of the seasonal variation in transmission and β0 > 0 is the time-averaged contact rate. This form of the seasonal contact rate captures the first mode of periodic change in disease transmission due to the school terms and holidays [39]. 2 2) The exposed individuals become infectious at the rate χ. Hence, 1/χ is the average latent period of the disease. 3) The infective individuals recover at the rate γ be- coming permanently immune. The average infectious pe- riod of the disease is 1/γ. 4) All individuals are subjected to the per capita death rate µ which is equal to the birth rate of the population. The average lifespan is given by 1/µ. The seasonally forced deterministic SEIR model can now be written as follows ds dt de dt di dt = µ (1 − s(t)) − β0(1 + β1 cos 2πt)s(t)i(t), = β0(1 + β1 cos 2πt)s(t)i(t) − (χ + µ)e(t), = χe(t) − (γ + µ)i(t), (2) (3) (4) where s(t), e(t) and i(t) denote the fraction of suscepti- ble, exposed and infective individuals, respectively. Note that the equation for the fraction of recovered individu- als, r(t), is redundant since s(t) + e(t) + i(t) + r(t) = 1, where the size of the total population was normalized to unity. The unforced deterministic SEIR model is recovered by setting β1 = 0 [1, 16]: ds dt de dt di dt = µ (1 − s(t)) − β0s(t)i(t), = β0s(t)i(t) − (χ + µ)e(t), = χe(t) − (γ + µ)i(t). (5) (6) (7) Both model (2)-(4) and model (5)-(7) have been ex- tensively investigated in the literature [16 -- 20]. Here, we will review well known facts about these models that are relevant for the present study. The dynamics of the un- forced deterministic SEIR model (5)-(7) depends on the basic reproductive ratio of the infection [1, 16] R0 = β0χ (χ + µ)(γ + µ) , (8) defined as the average number of secondary cases caused by an infectious individual in a totally susceptible pop- ulation in one infectious period. The stability analysis shows that Eqs. (5)-(7) have an asymptotically stable endemic equilibrium (s∗, e∗, i∗) = (cid:18) 1 R0 , µ(γ + µ)(R0 − 1) β0χ , µ(R0 − 1) β0 (cid:19) , provided R0 > 1. The trivial steady state (s∗, e∗, i∗) = (1, 0, 0) is stable for R0 < 1 and unstable for R0 > 1. (9) (10) Rate of disease onset Average latent period Fixed parameters χ 1/χ Rate of recovery Average infectious period Birth/death rate Average lifespan Average contact rate Basic reproductive ratio γ 1/γ µ 1/µ β0 R0 Variable parameter Amplitude of seasonal forcing β1 35.84 year−1 10.18 days 100 year−1 3.65 days 0.02 year−1 50 years 1575 year−1 15.74 0.02 (T = 1) 0.05 (T = 1) 0.10 (T = 1, 3) 0.12 (T = 2, 3) 0.2 (T = 2) TABLE I: Parameter values for measles that will be used in this paper. According to Eq. (8) this set corresponds to R0 = 15.74. The amplitude of seasonal forcing will be varied so that the solutions of Eqs. (2)-(4) exhibit stable limit cycles of period T indicated in the parenthesis next to the β1 value. If the contact rate varies seasonally according to Eq. (1) then a wide range of dynamical behavior of the sea- sonally forced deterministic SEIR model (2)-(4) is possi- ble, depending on R0 and on the amplitude of seasonal forcing β1. In what follows, we will restrict our analysis to the case of measles and use the parameter values given in Refs. [17, 19]. The birth/death rate will be fixed at 0.02 year−1 which corresponds to the average lifespan of 50 years. The average latent and infectious periods will be equal to 10.18 days and 3.65 days, respectively, yielding 13.83 days for the average time interval between infection and recovery. These values correspond to 35.84 year−1 for the disease onset rate and 100 year−1 for the recovery rate. Finally, the average contact rate is adjusted to be 1575 year−1 so that the basic reproductive ratio is 15.74. The remaining parameter, the forcing amplitude, will be given different values in the interval [0.02, 0.2] (estimates for β1 can be found in Ref. [14]). The summary of the parameter values is shown in Table I. The bifurcation analysis of the seasonally forced deter- ministic SEIR model (2)-(4) for the fixed set of parameter values considered above and β1 as a single free param- eter can be found in Refs. [18] an analysis of the same model was performed for variable β1 and R0 = 18 (corresponding to β0 = 1800 year−1). For the interested reader, we recommend Ref. [20] where more general bifurcation diagrams of Eqs. (2)-(4) were computed with two free parameters R0 and β1 and the remaining parameters held constant as in Table I. [17, 19]. In Ref. For the parameter values of Table I and variable β1, a brief summary of the bifurcation diagram is as follows [17, 19]. If β1 is positive but small, a stable limit cycle of 3 x 10−4 (a) β 1 2.7 2.4 = 0.05 − annual cycle i(t) 2 1.6 2 3 4 5 6 7 8 9 10 = 0.2 − biennial cycle 1 1.2 0 x 10−4 (b) β 1 8 i(t) 6 4 2 2 3 4 5 6 7 8 9 10 1 0 0 x 10−3 (c) β 1 2 2.5 = 0.1 − triennial cycle 1.5 i(t) 1 0.5 0 0 1 2 3 4 5 t 6 7 8 9 10 FIG. 1: The fraction of infective individuals as a function of time for periodic solutions of Eqs. (2)-(4). period 1 bifurcates from the endemic equilibrium point (9). As β1 is increased monotonically, it is found that at a value of β′ 1 ≈ 0.11479 a stable branch of period 2 bifur- cates from the period 1 branch, and the period 1 branch becomes unstable for β1 > β′ 1. Additionally, in some range of β1 these period 1 and period 2 limit cycles coex- ist with a pair of limit cycles of period 3 (one stable and one unstable) appearing from a saddle-node bifurcation. Finally, in a very narrow window of β1 the period 2 and period 3 branches exhibit a cascade of period-doubling bi- furcations as β1 increases, leading to chaotic epidemics. The full bifurcation diagram contains yet another stable attractor of period 4 that coexists with stable period 2 and period 3 limit cycles, however this attractor has a very small basin of attraction and it is hard to spot both in numerical integrations of Eqs. (2)-(4) and in stochastic simulations. Thus, in the realistic parameter region there are three main stable attractors, namely stable limit cycles of pe- riod 1, 2 and 3. We will consider typical values of β1 for which i) only a limit cycle of period 1 exists (β1 = 0.02, 0.05); ii) limit cycles of period 1 and 3 co- exist (β1 = 0.1); iii) limit cycles of period 2 and 3 coexist (β1 = 0.12); iv) a limit cycle of period 2 exists (β1 = 0.2). From now on the reader can refer to Table I where the periods T of the (co)existing limit cycles are indicated for each value of β1. Note that the regimes ii) and iii) lie in the vicinity of the bifurcation point where a stable limit cycle of period 2 is born (β′ 1 ≈ 0.11479). The fraction of infectives for stable periodic solutions of the seasonally forced SEIR model (2)-(4) is shown in Fig. 1. The solid curves were computed by numerical in- tegration of Eqs. (2)-(4) for the parameters values given in Table I and three values of the forcing amplitude. In all cases, initial conditions were chosen on a cycle so as to avoid transients. The annual epidemic cycle [β1 = 0.05, Fig. 1 (a)] corresponds to a small amplitude stable limit cycle of period 1 bifurcating from the endemic equilib- rium (9) of the unforced SEIR model (5)-(7). This cycle is present in the system if 0 < β1 < 0.11479. The equilib- rium value of infectives, i∗, in the unforced model (5)-(7) is also shown (the dashed line). The biennial epidemic cycle representing the alternating years of high and low incidence can be found for 0.11479 < β1 ≤ 0.2 (here we do not consider values of β1 > 0.2 where a stable limit cycle of period 2 is still present because such levels of forcing are considered unrealistically high). A plot of typical biennial epidemic is shown for β1 = 0.2 in Fig. 1 (b). Such a behavior of measles is in accordance with the data from the New York City and from England and Wales recorded between 1950 and 1968 (before vaccina- tion) [14, 16, 38]. As for other stable attractors coex- isting with stable limit cycles of period 1 and 2 these are, mainly, limit cycles of period 3 as mentioned in this section before. A typical limit cycle of period 3 is char- acterized by higher amplitude outbreaks and by very low minima of infective incidence, see Fig. 1 (c) where the behavior of the solution for the fraction of infectives is shown for β1 = 0.1. III. THE SEASONALLY FORCED STOCHASTIC EPIDEMIC MODEL In the seasonally forced stochastic SEIR model, a dis- crete population of the constant size N is divided into the classes of susceptibles (S), exposed (E), infectives (I) and recovered (R). At any instant of time, the total number of individuals equals N thus only three of the classes are independent. Denoting numbers of susceptible, exposed and infective individuals by m1, m2 and m3, respectively, and considering the processes postulated in Section II, we can obtain the transition rates for the stochastic model as follows. 1) The infection process takes place between a suscep- tible and an infective individual at the time dependent contact rate β(t) and results in the transition of the sus- −→ E + I. The ceptible to the exposed class, S + I corresponding transition rate is β(t) T m1−1,m2+1,m3 m1,m2,m3 = β0(1 + β1 cos 2πt) m1 N m3, (11) where the superscript and the subscript of T denote the final and the initial states of the system. 4 2) The transition of an individual from the exposed −→ I, occurs at the rate χ. class to the infective class, E The transition rate of this process equals χ T m1,m2−1,m3+1 m1,m2,m3 = χm2. (12) individual occurring at the rate γ, I 3) The transition rate of the recovery of an infective −→ R, is given by (13) = γm3. γ T m1,m2,m3−1 m1,m2,m3 The recovered individuals are permanently immune. µ with the linked birth, R S −→ R and I 4) Finally, there are four transition rates associated −→ S, and death processes, −→ R, occurring at the rate µ: (14) −→ R, E = µN, µ µ µ m1,m2,m3 m1,m2,m3 T m1+1,m2,m3 T m1−1,m2,m3 T m1,m2−1,m3 T ′m1,m2,m3−1 m1,m2,m3 m1,m2,m3 = µm1, = µm2, = µm3. (15) (16) (17) Note that the initial and final states for the recovery of an infective individual and the death of an infective in- dividual are equal, see Eqs. (13) and (17) but the cor- responding transition rates are different. To distinguish between the latter we use a prime in Eq. (17). The dynamics of the stochastic system determined by Eqs. (11)-(17) is completely described by the master equation [40, 41]. Given the initial and boundary condi- tions, this equation expresses the evolution of the prob- ability of having a system in state with m1 susceptibles, m2 exposed and m3 infectives, P(m, t), at any positive time. Denoting m as the shorthand for three numbers m1, m2, m3 this equation is written in the following form: dP(m, t) dt = Xm′6=mhT m m′P(m′, t) − T m′ m P(m, t)i. (18) The positive term on the r.h.s. of Eq. (18) is the increase of P(m, t) due to the transitions from states m′ to state m, while the negative term is the decrease due to the transitions from state m to states m′. The term with m equal to m′ is excluded because its contribution is zero. Substituting Eqs. (11)-(17) into Eq. (18) one obtains the master equation for the seasonally forced stochas- tic SEIR model (46) (see the Appendix). This equation is in fact a system of m1 · m2 · m3 coupled differential- difference equations similar to those used in studies of simple birth-death stochastic processes [41, 42]. In that case, an exact solution can be found in terms of generat- ing functions or by other means but Eq. (46) with three variables m1, m2 and m3 is too complicated to be solved exactly. To get insight into the dynamics of the season- ally forced stochastic SEIR model for large system sizes, we will apply a method due to van Kampen [40]. Sim- ilarly to other studies [32, 33], we will expand Eq. (46) in powers of the inverse of the system size N . First, let us rewrite Eq. (46) in terms of the step operators ǫ±1 , where j = 1, 2, 3, defined by their action on a smooth function f (m1, m2, m3, t) [40]: j where ǫ±1 1 f (m1, m2, m3, t) = f (m1 ± 1, m2, m3, t), ǫ±1 2 f (m1, m2, m3, t) = f (m1, m2 ± 1, m3, t), ǫ±1 3 f (m1, m2, m3, t) = f (m1, m2, m3 ± 1, t). (19) (20) (21) and Using Eqs. (19)-(21) the master Eq. (46) transforms into Eq. (47) given in the Appendix. For the seasonally forced stochastic SEIR model the expansion is made around a deterministic periodic solu- tion of Eqs. (2)-(4) which is a stable limit cycle of period T (see Ref. [33]). Thus, we make a transformation from the discrete variables mj(t) to the continuous variables xj(t), where j = 1, 2, 3, according to the equations m1(t) = N ¯s(t) + √N x1(t), m2(t) = N ¯e(t) + √N x2(t), m3(t) = N¯i(t) + √N x3(t), (22) (23) (24) where ¯s(t), ¯e(t) and ¯i(t) denote the deterministic trajec- tory of Eqs. (2)-(4) and x1(t), x2(t) and x3(t) are the corresponding stochastic fluctuations about it. After the substitution of Eqs. (11)-(17) and Eqs. (22)-(24) into Eq. (47), the terms of different orders in N can be identified in Eq. (47). The leading order terms yield Eqs. (2)-(4) with the substitutions s(t) = ¯s(t), e(t) = ¯e(t), i(t) = ¯i(t): d¯s dt d¯e dt d¯i dt = µ (1 − ¯s(t)) − β0(1 + β1 cos 2πt)¯s(t)¯i(t), = β0(1 + β1 cos 2πt)¯s(t)¯i(t) − (χ + µ)¯e(t), = χ¯e(t) − (γ + µ)¯i(t). (25) (26) (27) The next-to-leading order terms yield a multivariate lin- ear Fokker-Plank equation for the probability distribu- tion Π(x, t) [40, 41] ∂Π ∂t = −Xj,k Ajk(t) ∂(xkΠ) ∂xj + 1 2 Xj,k Bjk(t) ∂2Π ∂xj∂xk , (28) where j, k = 1, 2, 3. In Eq. (28), A(t) is the Jacobian of Eqs. (2)-(4) A(t) =   −µ − β(t)¯i(t) β(t)¯i(t) 0 0 −β(t)¯s(t) −(χ + µ) β(t)¯s(t) −(γ + µ) χ   (29) and B(t) is the symmetric cross correlation matrix f11(t) = µ(1 + ¯s(t)) + f12(t), f22(t) = µ¯e(t) + f12(t) + f23(t), f33(t) = (γ + µ)¯i(t) + f23(t) f12(t) = β(t)¯s(t)¯i(t), f23(t) = χ¯e(t). 5 (31) (32) (33) (34) (35) Both matrices A(t) and B(t) are evaluated on the limit cycle solutions of Eqs. (2)-(4) and thus they are periodic functions of t with the same period T of the limit cycle, A(t) = A(t + T ), B(t) = B(t + T ). (36) In previous studies of unforced epidemic models [28 -- 31], the power spectrum of stochastic fluctuations xj (t) was computed from the multivariate Langevin equation associated with Eq. (28) [40, 41] dxj(t) dt = Xk Ajk(t)xk(t) + Lj(t), j, k = 1, 2, 3, (37) where Lj(t) are Gaussian noise terms with zero mean hLj(t)i = 0 and with the correlator hLj(t)Lk(t′)i = Bjk(t)δ(t − t′). (38) (39) The analytical calculation of the power spectrum through the Fourier transform of Eq. (37) done in those stud- ies depends on the fact that the matrices A and B are constant in the unforced case. For the seasonally forced stochastic SEIR model, the matrices A(t) and B(t) are time dependent and this method does not work anymore. However, one can use the periodicity of A(t) and B(t) and Floquet's theory to find a solution of Eq. (37) and compute its power spectrum, as briefly outlined below. This method was developed to study the effects of ex- ternal noise in nonlinear oscillations close to bifurcations [43, 44], and it has been applied to intrinsic stochasticity for several systems similar to the model considered here [33]. The general solution of the inhomogeneous system of first-order linear differential equations (37) with matrix function A(t) and vector function L(t) can be written as a sum of the general solution of the homogeneous system dxj (t) dt = Xk Ajk(t)xk(t), j, k = 1, 2, 3, (40) B(t) =   f11(t) −f12(t) −f12(t) 0 0 f22(t) −f23(t) f33(t) −f23(t) ,   (30) and a particular solution of the inhomogeneous system [45]. Furthermore, the general solution of the homoge- neous system (40) with A(t) periodic with period T obeys Floquet's theorem [45, 46]. This theorem states that if X(t) is a fundamental matrix of Eq. (40), then there exists a periodic nonsingular matrix Q(t) with period T and a constant matrix R such that X(t) = Q(t)etR, Q(t) = Q(t + T ). (41) The matrix D = eT R is sometimes referred to as the monodromy matrix of the fundamental matrix X(t). Al- though D is not unique, its eigenvalues, λ1, λ2, λ3, called the characteristic (Floquet) multipliers associated with the periodic matrix A(t), are unique. The eigenval- ues of matrix R, ρ1, ρ2, ρ3, are called the characteristic (Floquet) exponents associated with the periodic matrix A(t). The latter are related to the characteristic multi- pliers by ρj = 1 T log λj, j = 1, 2, 3, (42) where the principal value of the logarithm is taken. Using further Floquet's theory an analytical expression can be obtained for the autocorrelation function of the stochastic fluctuations Cjj (τ ) = 1 T T Z 0 hxj (t)xj (t + τ )idt, j = 1, 2, 3, (43) in terms of the matrices Q(t), R and B(t) [33, 43]. Note that Eq. (43) does not depend on the initial condition if the initialization time is set to the infinite past. The power spectrum (power spectral density) Pj(ω) of the stochastic fluctuations can then be computed from the autocorrelation function Cjj (τ ) via the Fourier transform (the Wiener-Khinchin theorem) Pj(ω) = Cjj (τ )e−iωτ dτ. (44) +∞ Z −∞ In the following section, we will compare the power spectra calculated analytically through the method de- scribed above with those measured from simulations of the stochastic process defined by Eqs. (11)-(17). The spectra of the fluctuations of infectives are of particular interest to us because they can be compared with the spectra computed from real incidence data [22, 47]. IV. RESULTS The stochastic process is simulated with the use of a modified Gillespie algorithm [48] to account for the ex- plicit time dependence in the transition rates [49, 50]. Unless explicitly stated otherwise, all simulations start from a random initial condition and time series of 500 years or 100 years are recorded (50 or 10 years of which 6 β1 0.02 0.05 0.10 0.12 0.2 T 1 1 1 3 2 3 2 Floquet multipliers −0.8262 ± i 0.3007 Floquet exponents −0.1287 ± i 2.7926 9.1867 × 10−60 −135.9373 −0.8369 ± i 0.2696 −0.1287 ± i 2.8299 9.1868 × 10−60 −135.9373 −0.8724 ± i 0.1091 −0.1287 ± i 3.0172 9.1866 × 10−60 −135.9374 −0.2715 ± i 0.6119 −0.1338 ± i 0.6628 7.7134 × 10−178 0.7618 ± i 0.1310 8.4390 × 10−119 −135.9391 −0.1287 ± i 0.0852 −135.9374 −0.6562 ± i 0.1305 −0.1340 ± i 0.9878 7.7122 × 10−178 0.1184 ± i 0.7630 8.4357 × 10−119 −135.9391 −0.1293 ± i 0.7085 −135.9376 TABLE II: Floquet multipliers, λ1, λ2, λ3, and Floquet ex- ponents, ρ1, ρ2, ρ3, for limits cycles of different periods and several values of the forcing amplitude. All other parameters are kept fixed as in Table I. are considered as transient). By definition the model does not include injection of infectives, thus if an extinction occurs before this time a simulation is discarded. The largest system size we tested is N = 108 and the smallest one is N = 5×105. For some amplitudes of seasonal forc- ing the number of extinctions gets huge and we were not able to run such long simulations for system sizes smaller than N = 106. However, shorter time series could have been obtained (for comparison with the prevaccination era these should be about 20 years long [14]). There is one difficulty in the computation of the an- alytical autocorrelation functions and analytical power spectra. Although an explicit expression for the autocor- relation function can be found, the final results have to be obtained numerically because the stable limit cycle is not known in closed form. In the endemic equilibrium, the eigenvalues of the Jacobian of the unforced SEIR deter- ministic model (5)-(7) differ by two orders of magnitude for the parameter values typical of childhood infectious diseases. This is reflected in the seasonally forced deter- ministic SEIR model (2)-(4) [17] in the same parameter region. For the stable limit cycles given in Table I, two of the characteristic multipliers are always complex con- jugate and have real part of order 1, while the third one, λ3, is of order 10−59, 10−118 and 10−177 for limit cycles of periods 1, 2 and 3, respectively. In this way, the mon- odromy matrix D becomes singular (det D = λ1λ3λ3) in double precision numerical calculations and the periodic matrix Q(t) cannot be computed. One way to circumvent this difficulty is to reduce the computations to the two di- mensional central manifold associated with the complex eigenvalues, disregarding the dynamics along the strongly stable manifold associated with the small real eigenvalue λ3 where fluctuations will be strongly damped (a similar approach was used heuristically in Ref. [22]). Here we adopt a direct approach by implementing a Runge-Kutta 7-8 method for the numerical integration of the differen- tial equations on the limit cycle and by using arbitrary precision libraries NTL and GMP [51]. Typically the working precision set up in the integrations was 5 dig- its higher than the smallest characteristic multiplier (for example, 65 digits for limit cycles of period 1) and the numerical trajectory was correct up to 50 digits. Thus we could perform all the computations in the original variables of the full three dimensional system, which al- lows an immediate comparison with the power spectra of the simulations. The summary of the Floquet multipliers and Floquet exponents computed in this way is given in Table II. Note that until now we have not raised the question of the stability of the deterministic limit cycles because we do know from previous studies that the cycles are stable [17, 19, 20]. However, we automatically check this by computing the absolute values of λj which must be less than unity for a limit cycle to be asymptotically stable (or, alternatively, the real parts of ρj must be negative) [45, 46]. The homogeneous equation (40) is, in fact, a variational equation for small perturbations xj (t) around the periodic solution of Eqs. (2)-(4) [45, 46]. The pres- ence of two complex conjugate Floquet multipliers im- plies that deterministic perturbations decay to the cycle in a damped oscillatory way. This situation is similar to that of unforced systems with a stable focus in which resonant amplification of stochastic fluctuations was ob- served [27, 29]. It will become clear in what follows that in the stochastic system with seasonal forcing fluctua- tions around the limit cycles are significantly amplified too, and that the analytical and simulated power spectra cannot be explained by the deterministic theory alone. To validate the theory developed in this study, we first compare analytical and simulated autocorrelation func- tions for the stochastic fluctuations of infectives. Typical values of the forcing amplitude are chosen in the annual and biennial regime (see left and right upper plots of Fig. 2). Far from bifurcation points and for large system sizes, we find an excellent agreement in both cases (a small di- vergence can occur because of the sparse discretization of the orbit which we are lead to do to avoid very heavy computations). The lower plots of Fig. 2 show the corre- sponding power spectra. More exactly, the x-axes stands for temporal frequency and the y-axes is the power of the discrete Fourier transform of the autocorrelation func- tion time series. In this and in the following plots, the power spectra are normalized so that the total power is the summed squared amplitude of the time series [52]. One observes that the power spectra exhibit a number of peaks occurring regularly. As it has been noticed be- fore by several authors [33, 43], the peaks are located at 7 β 1 = 0.2 − biennial cycle β = 0.05 − annual cycle 1 0.015 0.01 0.005 0 −0.005 −0.01 10 τ 20 −0.015 0 30 5 10 τ 15 20 0.01 0.005 0 ) τ ( 3 3 C −0.005 −0.01 0 102 101 100 10−1 r e w o p 102 101 100 10−1 2 0 0 1 0.5 1.5 frequency (year−1) 1 0.5 1.5 frequency (year−1) 2 FIG. 2: (Color online) The upper panels show the autocor- relation functions of the stochastic fluctuations of infectives. The theoretical curves are plotted in gray (blue), and the curved computed from the simulations are plotted in black (for the annual cycle the analytical and simulated autocor- relation functions are strictly coincident). The lower panels show the corresponding power spectra. The vertical helper lines mark the frequencies predicted by Eq. (45). The dashes (dotted) lines are calculated by taking the plus (minus) in the equation. The system size used for simulation is 108. frequencies νn = n T ± Im ρ1,2 2π , (45) where n = 0,±1,±2, . . ., Im ρ1,2 denotes the absolute value of the imaginary part of complex conjugate Floquet exponents (see Table II), and T is period of a limit cycle. Note that for the annual limit cycle the main stochastic peak is situated at Im ρ1,2/(2π) while for the biennial cycle this peak is much smaller than the peak at 1/2 − Im ρ1,2/(2π) which is dominant. The plots are done in lin-log scale for a better observation of the structure of the power spectra. The analytical predictions for stochastic power spectra work well in a large range of β1 for annual and biennial limit cycles. However, as the system size is decreased sys- tematic deviations start to appear. We first consider the case of the annual limit cycle induced by a low forcing am- plitude. In Fig. 3 we compare the power spectra for the number of infective individuals from simulations for sev- eral system sizes N . In all panels, sharp peaks due to the annual limit cycle can be observed at integer frequencies, as well as several broader stochastic peaks whose frequen- cies are given by Eq. (45). For all system sizes the first (and largest) stochastic peak is situated at Im ρ1,2/(2π), r e w o p 105 104 103 102 r e w o p 105 104 103 102 1010 108 106 104 107 105 103 N=108 0.5 1 1.5 2 2.5 N=5×106 107 105 103 106 104 0.5 1 1.5 2 102 2.5 N=107 0.5 1 1.5 2 2.5 N=106 1 0.5 N=5×105 1.5 2 2.5 1010 108 106 104 109 107 105 103 N=108 0.5 1 1.5 2 N=5×106 0.5 1 1.5 2 109 107 105 103 2.5 106 104 102 2.5 8 N=107 0.5 1 1.5 2 2.5 N=106 1 0.5 N=5×105 1.5 2 2.5 1 0.5 2 frequency (year−1) 1.5 2.5 1 0.5 2 frequency (year−1) 1.5 2.5 FIG. 3: Power spectra of the number of infectives calculated from simulations for several system sizes N . The simulations of 500 years were run for N = 108, 107, 5 × 106, 106 and of 100 years for N = 5 × 105. Averages over 103 realizations were made to obtain each curve. The forcing amplitude is low and corresponds to the annual limit cycle, β1 = 0.02. The vertical helper lines mark the frequencies predicted by Eq. (45). the second is situated at 1−Im ρ1,2/(2π), etc. However, for N = 106 and N = 5×105 the dominant frequencies of the stochastic peaks are very slightly shifted to the left, so the characteristic period of the stochastic fluctuations gets higher. This effect has already been observed before in the spectra of fluctuations around a stable node, in unforced systems [28]. We also find that the spectra are fully described by the theory only for very large popula- tions. Beginning with N = 107 a number of much smaller stochastic peaks starts to appear close to the determin- istic peaks. Their amplitudes increase as the population size decreases but they stay orders of magnitudes lower than the dominant stochastic and deterministic peaks. These secondary peaks cannot be explained within the theory developed in this study and require considering corrections to the linear Fokker-Planck equation. An- other effect apparent in small systems is the change in the relative amplitude of the main stochastic and deter- ministic peaks. For small populations the deterministic limit cycle does not dominate the dynamics of the sys- tem any more. The enhancement and broadening of the stochastic peaks show a much noisier and irregular dy- namics. For comparison, see the panel for N = 5 × 105 in Fig. 3 where the main stochastic peak is significantly FIG. 4: Power spectra of the number of infectives calculated from simulations for several system sizes N . The simulations of 500 years were run for N = 108, 107, 5 × 106, 106 and of 100 years for N = 5 × 105. Averages over 104 (103) realizations were made to obtain the power spectra for N = 108 (for the remaining system sizes). The forcing amplitude is 2.5 higher than in Fig. 3 but the deterministic system is still in the annual limit cycle regime, β1 = 0.05. The vertical helper lines mark the frequencies predicted by Eq. (45). higher and broader than the main deterministic peak. The same picture of changes in the spectra for different population sizes is maintained if the amplitude of forcing is increased but the annual limit cycle is still stable. As an example, see Fig. 4 done for β1 = 0.05 (this forcing amplitude is 2.5 larger than the one used in Fig. 3). If the forcing amplitude is increased even further the period doubling of the limit cycle occurs at β′ 1 ≈ 0.11479 and everywhere in the interval 0.11479 < β1 ≤ 0.2 a sta- ble limit cycle of period 2 is present in the deterministic model. In Fig. 5 we compare the spectra from simu- lations and the analytical spectra for a range of system sizes. Again, the spectra demonstrate narrow determin- istic peaks at multiples of 1/2 due to the limit cycle of period 2 and regular stochastic peaks. For the largest system size, the stochastic peaks are predicted correctly by the theory. The major stochastic peak is now lo- cated at 1/2 − Im ρ1,2/(2π). Note that the peak which used to be dominant in the annual regime was located at Im ρ1,2/(2π). But now this frequency corresponds to the smallest among all stochastic peaks in the range of frequencies shown in the plot. This observation de- r e w o p r e w o p 1011 109 107 105 109 107 105 103 N=108 1010 108 106 N=107 0.5 1 1.5 2 104 2.5 0.5 1 1.5 2 2.5 N=5×106 1 0.5 2 frequency (year−1) 1.5 107 105 103 2.5 N=106 1 0.5 2 frequency (year−1) 1.5 2.5 FIG. 5: Power spectra of the number of infectives calculated from simulations for several system sizes N . The simulations of 500 years were run for N = 108, 107, 5 × 106 and of 100 years for N = 106. Averages over 5 × 103 (103) realizations were made to obtain the power spectra for N = 108 (for the remaining system sizes). The forcing amplitude corresponds to the biennial limit cycle, β1 = 0.2. The vertical helper lines mark the frequencies predicted by Eq. (45). serves a longer comment and we will go back to it later. For smaller values of N we observe that the determin- istic peak at 1/2 gets smaller and merges with the two surrounding stochastic peaks. At some point it becomes impossible to distinguish between the stochastic and de- terministic peaks which are represented by a single broad peak around 1/2. The stochastic peaks get significantly enhanced but also the range of frequencies present in the infective time series. The next parameter range of interest is close to the point where the period doubling bifurcation occurs for the deterministic system. In this regime we do not expect to have an agreement between the theory and simulations because none of the cycles is stable enough, however there are interesting conclusions which can be drawn from the power spectra. Recall that the period doubling in the deterministic model occurs at β′ 1 ≈ 0.11479 but in the stochastic model the transition from the annual limit cy- cle to the biennial limit cycle is shifted to a higher value of β1. In Fig. 6 we compare the spectra for the number of infectives from simulations performed for two close values of β1 in the vicinity of the deterministic period doubling point, one before the bifurcation (β1 = 0.1, correspond- ing to the stable annual limit cycle, gray (blue) curves) and one after (β1 = 0.12, corresponding to the stable biennial limit cycle, black curves). As one can see the transition is blurred and shows up later in the stochas- tic system. The same behavior of the fluctuations power spectrum around a bifurcation was described in Ref. [43]. 1011 109 107 105 109 107 105 103 N=108 0.5 1 1.5 2 N=5×106 1010 108 106 104 2.5 108 106 104 102 0.5 1 1.5 2 2.5 9 N=107 0.5 1 1.5 2 2.5 N=106 0.5 1 1.5 2 2.5 N=5×105 r e w o p 106 104 102 1 0.5 2 frequency (year−1) 1.5 2.5 FIG. 6: (Color online) Power spectra of the number of in- fectives calculated from simulations for several system sizes N . The gray (blue) and black curves were obtained for In the determinis- β1 = 0.1 and β1 = 0.12, respectively. tic model the period doubling occurs at β ′ 1 ≈ 0.11479. The frequencies of the vertical helper lines correspond to the an- nual limit cycle. The simulations of 500 years were used for N = 108, 107, 5 × 106, 106 and of 100 years for N = 5 × 105. Averages over 103 realizations were made to obtain all curves. 1013 1010 107107 r e w o p 1010 N=108 N=107 0.5 1 frequency (year−1) 105 1.5 1 0.5 2 frequency (year−1) 1.5 2.5 FIG. 7: (Color online) Power spectra of the number of infec- tives calculated from simulations for two system sizes N and two sets of initial conditions [for the gray (blue) curve sim- ulations started from random initial conditions and for the black curves the initial conditions were chosen close to the deterministic triennial cycle]. In the left (right) panel the vertical helper lines mark the predicted peak frequencies for the triennial (annual) limit cycle. The amplitude of seasonal forcing corresponds to coexisting stable annual and triennial limit cycles in the deterministic model, β1 = 0.1. There is almost no difference between the simulated spec- tra. The first deterministic peak signaling the biennial limit cycle has to be present at 1/2 for the black curves, and this is where the stochastic peaks for the annual limit cycle are located. The two main stochastic peaks of the annual cycle move on to 1/2 and at some point give rise a deterministic peak of the biennial cycle. This is seen from the plot for the largest system size, however the de- terministic peak becomes obvious if the β1 is increased a bit further. A similar broadening and shifting of the bifurcation point in simulations has also been observed in the unforced model with a transition from a stable focus to a stable limit cycle through an Andronov-Hopf bifurcation [29, 32]. The last question one naturally poses is how does the coexistence of attractors influence the power spectrum, through possible switches between basins of attraction induced by stochasticity. The relevance of this mecha- nism has been argued for in the literature [53, 54], in particular to try to provide an explanation for the di- versity of patterns of childhood diseases [21, 22]. In the deterministic model there is a range of β1 where stable annual and biennial limit cycles coexist with a stable tri- ennial cycle. In particular, this happens for the parame- ter values of Fig. 6. We, however, have not identified any deterministic peaks at multiples of 1/3 associated with a triennial limit cycle or even an indication of such a be- havior in the stochastic system. The same was confirmed for other values of β1 where the orbits coexist. To clar- ify this situation we performed an experiment in which one set of simulations started on a triennial limit cycle (these conditions are favorable for the triennial cycle as it will become clear later) and the other set started from a random initial condition. The resulting power spectra are shown in Fig. 7 for β1 = 0.1. The black [respectively, gray (blue)] curves are the power spectra of the infec- tive time series beginning from favorable (respectively, random) conditions. It is only for system size N = 108 that we were able to observe a triennial limit cycle with the fluctuations around it described by the van Kampen expansion (see the left panel in Fig. 7, the helper lines mark the frequencies for the triennial limit cycle). For all other system sizes the power spectra are identical to those already shown in Fig. 6. In this case, a better insight into what happens in the simulations is given by inspecting the time series. In Fig. 8 we show the density of infectives in a typical single realization of the stochastic model starting from the tri- ennial limit cycle (for the same value of β1 as in Fig. 7). For system size N = 108 the density exhibits reg- ular triennial oscillations. Their characteristic features are a high amplitude (at least twice as large as that of a typical annual or biennial cycle) and very low number of infectives between recurrent epidemics. For N = 108 the system tends to stay on the triennial limit cycle because the fluctuations are not large enough to drive it to the x 10−3 2 1.5 10 N=108 N / m 3 1 0.5 0 0 x 10−3 2 1.5 5 10 15 20 25 30 35 40 45 50 N=107 N / m 3 1 0.5 0 0 x 10−3 1 0.8 0.6 5 10 15 20 25 30 35 40 45 50 N=106 0.4 0.2 0 0 5 10 15 20 25 t 30 35 40 45 50 N / m 3 FIG. 8: The infective density recorded from a typical realiza- tion of the stochastic model starting on the deterministic tri- ennial limit cycle. The seasonal forcing amplitude is β1 = 0.1. N=108 5 10 15 20 25 30 35 40 45 50 N=107 5 10 15 20 25 30 35 40 45 50 N=5×106 N / m 3 N / m 3 x 10−4 4 3 2 1 0 0 x 10−4 6 4 2 0 0 x 10−3 1 0.8 0.6 N / 3 m 0.4 0.2 0 0 5 10 15 20 25 t 30 35 40 45 50 FIG. 9: The infective density recorded from a typical real- ization of the stochastic model starting on the deterministic annual limit cycle. The seasonal forcing amplitude is β1 = 0.1. N = 108 N = 107 5 × 105 ≤ N ≤ 106 11 β1 0.05 well defined dominant annual deterministic peak well defined dominant annual and deterministic peaks biennial 0.2 well defined dominant annual deterministic peak and two subdominant close to biennial broad stochastic peaks well defined dominant annual deterministic peak and subdom- inant broad biennial stochastic peak well defined annual determinis- tic peak and close to biennial broad stochastic peak well defined dominant annual deterministic peak and subdom- inant close to biennial much broader stochastic peak TABLE III: Qualitative features of the power spectra for different values of β1 and N . annual limit cycle. But for smaller N the system either goes extinct or drops very quickly to the annual limit cy- cle (or a biennial one if β1 > 0.11479) and stays on it. For the population of one million of individuals not even one oscillation over the triennial cycle is followed by the stochastic system because the relative fluctuations are large at the first drop of infectives and drive the system to the basin of attraction of the annual cycle. In Fig. 9 we show some of the time series for the same parameter values as in the previous figure but for initial conditions starting on the annual limit cycle. The density exhibits alternating regions of annual and biennial oscillations for large population sizes with an ever increasing role of the stochastic fluctuations as N decreases. For small sizes the time series look quite irregular, occasionally exhibit- ing interepidemic intervals longer than 2 years. This case corresponds to the power spectra where the stochastic peaks are broad and high. It is remarkable that we have never observed the backward switching from an annual (or biennial) limit cycle to a triennial one. The qualitative features of the power spectra for two different values of the forcing amplitude, β1, and different population sizes, N , are summarized in Table III. As dis- cussed above, well defined high amplitude peaks show up in the time series as regular cycles, and broad stochastic peaks as lower amplitude noisy oscillations. V. DISCUSSION AND CONCLUSIONS In conclusion, in this study we have considered the deterministic and the stochastic SEIR models with sinu- soidally varying contact rate. Depending on the forcing amplitude, the deterministic model exhibits a range of stable attractors, the most visible of which are limit cy- cles of periods 1, 2 and 3. Using the van Kampen's expan- sion of the master equation of the corresponding stochas- tic model we have calculated autocorrelation functions and power spectra of the stochastic fluctuations around these attractors. We have compared the analytical re- sults with those obtained from direct simulations of the stochastic model. We have found that in a large range of the forcing amplitude there is an excellent agreement be- tween the theory and simulations. The prerequisites for this are large system sizes (typically higher than 106) and the stability of attractors, namely the more stable the cycle and the higher the system size are the better the agreement between the simulated and analytical power spectra is. This is exactly what we would expect for the quality of the approximation given by the truncation at first order of the full van Kampen expansion. The power spectra of the infective time series demon- strate peaks of two types. The narrow peaks are due to a limit cycle of a given period and the broader peaks are due to the resonant amplification of stochastic fluctua- tions around the limit cycle. It has been argued that the presence and position of deterministic and stochas- tic peaks in the power spectra obtained from real data records of childhood diseases can be predicted by the deterministic theory alone, and that, moreover, the fre- quency of the stochastic peak is defined by the frequency of the transients near a stable limit cycle [22, 23, 47]. We have identified the main frequencies of the stochas- tic peaks in the annual and biennial regimes and shown that these do not necessarily equal the frequency of the damped oscillations of deterministic perturbations around a cycle. Thus, neither the full structure of the power spectrum of the infective time series nor the most prominent frequency of the stochastic peak can be fully predicted by the deterministic model. Another conclusion concerns the role of the coexis- tence of attractors in the seasonally forced stochastic and deterministic SEIR models. The coexistent stable limit cycles present in the deterministic epidemic mod- els have been conjectured to be the reason why irregu- lar dynamics is observed in the corresponding stochastic system. Namely, it has been a systematic assumption of several papers on childhood infectious diseases mod- eling [19, 21, 22] based exclusively on the deterministic analysis that for small populations the stochastic system should constantly switch between cycles of different pe- riods due to demographic or environmental noise, thus exhibiting irregular dynamics. The results of the present study contradict this view. For values of β1 close to the period doubling bifurcation value, demographic as well as environmental noise can promote switching between pe- riod 1 and period 2 cycles. However, for the stable limit cycles of period 1 or 2 coexisting with a stable cycle of pe- riod 3 such switching is absent in the stochastic system. Notwithstanding the basin of attraction of the triennial limit cycle being roughly 25% of the total initial condi- tions [19] for the deterministic system, it is extremely unstable in the simulations. Even for large system sizes the system is brought to a cycle of lower period after a very short time, and backwards switching is not observed. This can be understood from the shape of the periodic solutions of the triennial cycle of the deterministic model, together with the fact that most initial conditions with very low number of infectives are attracted to the lower period cycles. We have become aware of a recent paper where the fluc- tuation power spectrum of a seasonally forced stochastic SIR model is computed through the same method [55]. In this paper, a transition representing injection of in- fectives is included in addition to the processes of infec- tion, recovery, and birth-death. It is shown that, even for very low infective immigration rates, the inclusion of this term has a drastic impact on the bifurcation diagram of the deterministic model, leaving stable annual and bi- ennial limit cycles as the only possible attractors [55]. This shows that the competing higher period attractors of the forced system are very fragile indeed. They are not robust, in the deterministic setting, with regard to small changes in the model. On the other hand, in the stochastic setting, their effective basin of attraction is negligibly small, except for unrealistically large popula- 12 tion sizes. The results of this study and those of Ref. [55] concur to support the view that the main ingredient to understand the observed incidence patterns of childhood infectious diseases is stochastic amplification, rather than noise induced switching between competing attractors of the deterministic system. Acknowledgments Financial support from the Foundation of the Univer- sity of Lisbon (FUL) and the Portuguese Foundation for Science and Technology (FCT) under Contract No. POCTI/ISFL/2/261 is gratefully acknowledged. The first author (G.R.) was also supported by FCT under Grant No. SFRH/BD/32164/2006 and by Calouste Gul- benkian Foundation under its Program "Stimulus for Re- search". APPENDIX As described in the Section III of the main text, the master equation for the seasonally forced stochastic SEIR model is constructed by considering all possible transi- tions increasing or decreasing the probability of finding a system in state with m1 susceptible, m2 exposed and m3 infective individuals. This equation reads as follows: dP(m1, m2, m3, t) dt m1+1,m2−1,m3P(m1 + 1, m2 − 1, m3, t) + T m1,m2,m3 m1,m2+1,m3−1P(m1, m2 + 1, m3 − 1, t) + T m1,m2,m3 m1+1,m2,m3P(m1 + 1, m2, m3, t) + T m1,m2,m3 m1,m2,m3+1P(m1, m2, m3 + 1, t) −hT m1−1,m2+1,m3 m1,m2,m3 m1,m2,m3+1P(m1, m2, m3 + 1, t) m1−1,m2,m3P(m1 − 1, m2, m3, t) m1,m2+1,m3P(m1, m2 + 1, m3, t) + T m1,m2−1,m3+1 m1,m2,m3 = T m1,m2,m3 + T m1,m2,m3 + T m1,m2,m3 + T ′m1,m2,m3 + T m1,m2,m3−1 + T ′m1,m2,m3−1 m1,m2,m3 m1,m2,m3 m1,m2,m3 + T m1+1,m2,m3 iP(m1, m2, m3, t). + T m1−1,m2,m3 m1,m2,m3 + T m1,m2−1,m3 m1,m2,m3 (46) The above equation and its subsequent analysis can be simplified with the use of the step operators defined in the main text by Eqs. (19)-(21). The substitution of Eqs. (19)-(21) into Eq. (46) gives dP(m1, m2, m3, t) dt m1,m2,m3 3 − 1(cid:1)T m1,m2−1,m3+1 2 − 1(cid:1)T m1−1,m2+1,m3 = h(cid:0)ǫ2ǫ−1 + (cid:0)ǫ1ǫ−1 + (ǫ3 − 1)T ′m1,m2,m3−1 m1,m2,m3 m1,m2,m3 iP(m1, m2, m3, t). m1,m2,m3 + (ǫ1 − 1)T m1−1,m2,m3 +(cid:0)ǫ−1 1 − 1(cid:1)T m1+1,m2,m3 m1,m2,m3 m1,m2,m3 + (ǫ3 − 1)T m1,m2,m3−1 + (ǫ2 − 1)T m1,m2−1,m3 m1,m2,m3 (47) Press, Princeton, 2007). REFERENCES [1] M. J. Keeling and P. Rohani, Modeling Infectious Dis- eases in Humans and Animals (Princeton University 13 [2] R. Pastor-Satorras and A. Vespignani, Phys. Rev. Lett. [31] A. J. Black, A. J. McKane, A. Nunes and A. Parisi, Phys. 86, 3200 (2001). [3] M. E. J. Newman, Phys. Rev. E 66, 016128 (2002). [4] M. Girvan, D. S. Callaway, M. E. J. Newman, and Rev. E 80, 021922 (2009). [32] R. P. Boland, T. Galla, and A. J. McKane, J. Stat. Mech., P09001 (2008). S. H. Strogatz, Phys. Rev. E 65, 031915 (2002). [33] R. P. Boland, T. Galla, and A. J. McKane, Phys. Rev. E [5] J. Dushoff, J. B. Plotkin, S. A. Levin, and D. J. D. Earn, 79, 051131 (2009). Proc. Natl. Acad. Sci. U.S.A. 101, 16915 (2004). [34] H. W. Hethcote and P. van den Driessche, J. Math. Biol. [6] J. Verdasca, M. M. T. Gama, A. Nunes, N. R. Bernardino, J. M. Pacheco, and M. C. Gomes, J. Theor. Biol. 233, 553 (2005). [7] M. F. Boni, J. R. Gog, V. Andreasen, and W. Feldman, Proc. R. Soc., Ser. B 273, 1307 (2006). [8] J. C. Miller, Phys. Rev. E 76, 010101(R) (2007). [9] M. Marder, Phys. Rev. E 75, 066103 (2007). [10] G. Yan, Z.-Q. Fu, J. Ren, and W.-X. Wang, Phys. Rev. E 75, 016108 (2007). [11] L. B. L. Santos, M. C. Costa, S. T. R. Pinho, R. F. S. An- drade, F. R. Barreto, M. G. Teixeira, and M. L. Barreto, Phys. Rev. E 80, 016102 (2009). [12] V. R. V. Assis and M. Copelli, Phys. Rev. E 80, 061105 (2009). [13] V. Nagy, Phys. Rev. E 79, 066105 (2009). [14] M. J. Keeling and B. T. Grenfell, Proc. R. Soc. London, Ser. B 269, 335 (2002). [15] H. T. H. Nguyen and P. Rohani, J. R. Soc., Interface 5, 403 (2008). [16] R. M. Anderson and R. M. May, Infectious Diseases of Humans: Dynamics and Control (Oxford University Press, Oxford, 1991). [17] I. B. Schwartz and H. L. Smith, J. Math. Biol. 18, 233 (1983). [18] J. L. Aron and I. B. Schwartz, J. Theor. Biol. 110, 665 (1984). [19] I. B. Schwartz, J. Math. Biol. 21, 347 (1985). [20] Y. A. Kuznetsov and C. Piccardi, J. Math. Biol. 32, 109 (1994). [21] D. J. D. Earn, P. Rohani, B. M. Bolker, and B. T. Gren- fell, Science 287, 667 (2000). [22] C. T. Bauch and D. J. D. Earn, Proc. R. Soc. London, Ser. B 270, 1573 (2003). [23] P. Rohani, M. J. Keeling, and B. T. Grenfell, Am. Nat. 159, 469 (2002). [24] D. Alonso, A. J. McKane, and M. Pascual, J. R. Soc., Interface 4, 575 (2007). [25] R. Kuske, L. F. Gordillo, and P. Greenwood, J. Theor. Biol. 245, 459 (2007). [26] S. Azaele, A. Maritan, E. Bertuzzo, I. Rodriguez-Iturbe, and A. Rinaldo, Phys. Rev. E 81, 051901 (2010). [27] A. J. McKane and T. J. Newman, Phys. Rev. Lett. 94, 218102 (2005). 29, 271 (1991). [35] S. Ruan and W. Wang, J. Differential Equations 188, 135 (2003). [36] W. R. Derrick and P. van den Driessche, Discr. Cont. Dyn. Syst., Ser. B 3, 299 (2003). [37] E. Shim, Math. Biosci. Eng. 3, 557 (2006). [38] D. Schenzle, Inst. Math. Appl. J. Math. Appl. Med. Biol. 1, 169 (1984). [39] For the purposes of fitting to measles incidence time series, the term-time forced SEIR model is sometimes preferred. The sinusoidally forced SEIR model, however, reproduces the behavior of the term-time forced SEIR model at a lower level of forcing (see the comment (10) in Ref. [21]). Here we use Eq. (1) as it will somewhat sim- plify the analysis of stochastic effects in the SEIR model with a periodically varying contact rate still allowing us to understand the main mechanisms at play. [40] N. G. van Kampen, Stochastic Processes in Physics and Chemistry (Elsevier, Amsterdam, 1981). [41] H. Risken, The Fokker-Planck Equation (Springer, Berlin, 1996). [42] N. S. Goel and N. Richter-Dyn, Stochastic Models in Bi- ology (Academic Press, New York, 1974). [43] K. Wiesenfeld, J. Stat. Phys. 38, 1071 (1985). [44] K. Wiesenfeld, Phys. Rev. A 32, 1744 (1985). [45] R. Grimshaw, Nonlinear Ordinary Differential Equations (CRC Press, New York, 1993). [46] L. Meirovitch, Methods of Analytical Dynamics (McGraw-Hill, New York, 1970). [47] C. T. Bauch, in Mathematical Epidemiology, edited by F. Brauer, P. van den Driessche, and J. Wu (Springer, Berlin, 2008), p. 297. [48] D. T. Gillespie, J. Comput. Phys. 22, 403 (1976). [49] D. F. Anderson, J. Chem. Phys. 127, 214107 (2007). [50] T. Lu, D. Volfson, L. Tsimring, and J. Hasty, Syst. Biol. 1, 121 (2004). [51] NTL and GMP libraries http://www.shoup.net/ntl/ respectively. and available are at http://gmplib.org/, [52] W. H. Press, S. A. Teukolsky, and W. T. Vetterling, Nu- merical Recipes: The Art of Scientific Computing (Cam- bridge University Press, Cambridge, 2007). [53] M. J. Keeling, P. Rohani, and B. T. Grenfell, Physica D [28] M. Simoes, M. M. Telo da Gama, and A. Nunes, J. R. 148, 317 (2001). Soc., Interface 5, 555 (2008). [54] I. B. Schwartz, L. Billings, and E. M. Bollt, Phys. Rev. [29] G. Rozhnova and A. Nunes, Phys. Rev. E 79, 041922 E 70, 046220 (2004). (2009). [30] G. Rozhnova and A. Nunes, Phys. Rev. E 80, 051915 (2009). [55] A. J. Black and A. J. McKane, J. Theor. Biol. 267, 85 (2010).
1511.07177
2
1511
2016-05-06T08:26:52
Aquaporin-1 can work as a Maxwell's Demon in the Body
[ "physics.bio-ph" ]
Aquaporin-1 (AQP1) is a membrane protein which is selectively permeable to water. Due to its dumbbell shape, AQP1 can sense the size information of solute molecules in osmosis. At the cost of consuming this information, AQP1 can move water against its chemical potential gradient: it is able to work as one kind of Maxwell's Demon. This effect was detected quantitatively by measuring the water osmosis of mice red blood cells. This ability may protect the red blood cells from the eryptosis elicited by osmotic shock when they move in the kidney, where a large gradient of urea is required for the urine concentrating mechanism. This finding anticipates a new beginning of inquiries into the complicated relationships among mass, energy and information in bio-systems.
physics.bio-ph
physics
Aquaporin-1 can work as a Maxwell's Demon in the Body Liangsuo Shu1*, Yingjie Li*2, Xiaokang, Liu1Xin Qian1, Suyi Huang1, Shiping Jin1, and Baoxue Yang2 1Innovation Institute, School of Energy and Power Engineering, Huazhong University of Science & Technology. 1037 Luoyu Road, Wuhan, China; 2State Key Laboratory of Natural and Biomimetic Drugs, Department of Pharmacology, School of Basic Medical Sciences, Peking University, Beijing, China; *contributed to this work equally Xin Qian is at Mechanical Engineering, University of Colorado Boulder now. Address correspondence to: Shiping Jin Innovation Institute, School of Energy and Power Engineering, Huazhong University of Science & Technology. 1037 Luoyu Road, Wuhan, 430074, China; E-mail: [email protected] Tel: 0086-027- 87542618 Baoxue Yang Department of Pharmacology, School of Basic Medical Sciences, Peking University, 38 Xueyuan Road, Haidian District, Beijing, 100191, China. E-mail: [email protected] Fax: 0086-010-82805622 Key words: Maxwell's Demon, Aquaporin-1, Nano-osmosis Abstract Aquaporin-1 (AQP1) is a membrane protein which is selectively permeable to water. Due to its dumbbell shape, AQP1 can sense the size information of solute molecules in osmosis. At the cost of consuming this information, AQP1 can move water against its chemical potential gradient: it is able to work as one kind of Maxwell's Demon. This effect was detected quantitatively by measuring the water osmosis of mice red blood cells. This ability may protect the red blood cells from the eryptosis elicited by osmotic shock when they move in the kidney, where a large gradient of urea is required for the urine concentrating mechanism. This finding anticipates a new beginning of inquiries into the complicated relationships among mass, energy and information in bio-systems. Since 1867, Maxwell's Demon has been one of the focuses of thermodynamic debates and a hot topic of popular science1. After a series of theoretical works by many researchers, the demon had been proved to be able to convert information into free energy without contradicting the second law of thermodynamics2. In fact, some "demons" have been created in different ingenious experiments3–5 and more theoretical models to implement Maxwell's Demon have been proposed6–11. However, it is necessary to determine whether any Maxwell's Demons, which can convert information into free energy, exist in natural world and their scientific significance. Osmosis plays an important role in areas such as regulating water balance across cellular membranes12, the immune responses to pathogens of lymphocyte perforin13 and the membrane attack complex/perforin 14, and the function of pore-forming toxins15. In thermodynamics, it is generally accepted that osmotic pressure is the result of the chemical potential difference of the solvent across a semipermeable membrane, but the debate in osmosis dynamics never ends16. The discovery17 and follow-up studies18,19 of a nature nanoscale channel, aquaporin (AQP) give us a good opportunity for a better understanding of osmosis20. The researches of AQP osmosis reveal an unexpected phenomenon: the reflection coefficients of small impermeable solutes to AQP1 are smaller than 1 and have a close relation with their molecular sizes 21–23(Solomon-Hill effect). In a previous work, we got the quantitative explanation of this effect using an analytical method based on molecular dynamics20. The dumbbell shape of AQP1 makes it able to sense the size information of solute molecules in osmosis. However, this ability enables AQP1 to move water molecules against their chemical potential gradient in some cases. Considering the information consumption during the osmosis in these cases, we find Aquaporin-1 can work as a Maxwell's Demon which can convert information into free energy. The reflection coefficients of impermeable solutes (urea, mannitol, glucose) to AQP1 were measured using red blood cells(RBCs) lacking urea transporter-B (UT-B), from UT-B knockout mice to rule out the possible influences of the permeability of urea through UT-B24,25. Results reflection coefficients of impermeable solutes The osmotic pressure of a dilute solution is usually described by the famous Van't Hoff equation (1) where π is the osmotic pressure, c is the molar concentration of the solute, R is the molar gas constant, T is the thermodynamic temperature, and σ is the reflection coefficient. σ, introduced as a phenomenological coefficient26, was defined as the fraction of a certain solute that does not permeate the membrane, its retention. This definition has resulted in an inference that σ of a completely impermeable solute must be 1. However, many cRT experiments about AQP1 clearly indicate that σ values of some impermeable solutes are smaller than 1 and have a close relation to their molecular size21–23, the Solomon- Hill effect. It was found that this effect is a result of the special dumbbell-like shape of AQP120,22. As a remarked characteristic of all AQPs, the dumbbell-like shape is consists of an extracellular vestibule, a cytoplasmic vestibule, and a narrow channel connecting them 18. The vestibule surrounded by loops allow the incursions of solute molecules to contribute part of momentums in varying degrees depended on their molecular sizes20,22. In the channel of AQP, there is usually at least one filter which determines its selectivity. For AQP1, the filter is too narrow to allow any other molecules beside water to pass through. AQP1 contributes the main water permeability of RBCs and makes it an ideal materiel to measure the osmosis. In the experiments of Toon and Solomon23, they used common RBCs, which also expressed UT-B, as the materials to measure the reflection coefficient of solutes. This can lead to some degree of errors. On one hand, besides AQP1, UT-B and its plasma membrane also make some contribution of the water permeability of RBCs24,25,27. On the other hand, little solutes, such as urea, can permeate through UT-B and reduce the measured value of reflection coefficients. Therefore, Solomon-Hill effect was also suspected to a confounding effect of rapid diffusional urea transport in some literatures 24. To rule out the influence UT-B, we used UT-B null (UT-B-/-) RBCs to measure the reflection coefficients of solutes for AQP1 with stopped-flow. σs of solutes are calculated by comparing the initial rates of red cell volume changes in different solutions23,28,29. The dead times of our experiments were about 200 millisecond. Therefore, the slopes from 200 millisecond to 300 millisecond were used. For UT-B null RBCs (urea can be regarded as one impermeable solute), a widely used method of exponential function fitting 25,30–32 in reconstituted aquaporin osmotic permeability was also applied to calculate the reflection coefficients. The results calculated from two methods agreed well with each other. Fig. 1. Water permeability and reflection coefficients of AQP1 with different solutes in RBCs. Osmotic water permeability was measured from the time dependence of erythrocyte volume, in response to a 125-mM inwardly directed solute gradient, by stopped-flow light scattering. A, Representative curves showing water permeability in RBCs from wild-type mice (+/+) and UT-B null mice (-/-) measured at 37 °C under urea, mannitol or glucose and using a 125 mM gradient. B, averaged reflection coefficients of AQP1 with different solutes for experiments conducted as in A (mean ± S.E. n≥3 ) Our experimental results showed that the reflection coefficients of solutes for AQP1 increase with their molecular sizes (Fig.1). It was found that rapid diffusional urea transport through UT-B does make a difference to the reflection coefficients of urea [28], while no remarked difference was found for mannitol and glucose. However, this influence is rather limited compared with the size effect. For RBCs of UT-B null mice, σurea is reduced to 0.39 from 1 because of the size effect (the reflection coefficients of glucose is regarded as 1, σglucose=1). For erythrocytes of wild mice, σurea only reduces 0.09 to 0.30 because of the urea transport through UT-B. This can be attributed to the high water permeability of AQP. Both water osmosis through AQP1 and urea transporting through UT-B are non-equilibrium processes, during which time is an important factor. From the works of Zhao33, it can be found that the osmotically induced water efflux of red blood cells is much faster than urea transporting through UT-B. Therefore, the difference in response times of the two processes isolates them to a great extent. By ruling out the influence of UT-B, our improved experiments provided more accurate measurements of Solomon-Hill effect. Thermodynamics of Maxwell's Demon in Our Body The small reflection coefficients of impermeable solutes have also brought us a new question. Suppose the simplest of conditions: there is an erythrocyte and the solutes for the intracellular and extracellular solutions are pure glucose and pure urea respectively. When σureacurea< cglucose< curea, water is moved into the red blood cell through AQP1 against its chemical potential gradient. It is an entropy decreasing process which seems to break the second law of thermodynamics (Fig.2a). The entropy decrease(ds1-2) can be calculated as below, (2) where μex and μin are the chemical potential of extracellular and intracellular water respectively, dn and dV are the transferring amount and volume of water through AQP1 respectively. However, the real osmotic pressure difference sensed by the AQP1 in its osmosis is (3) and the entropy change during osmosis from the view of AQP1 is 12()()0exinglucoseureadndsRccdVT()glucoseureaureaRTcc Therefore, for AQP1, the osmosis is a spontaneous process of entropy increment, during which the second law of thermodynamics is strictly followed(Fig.2b). (4) Fig. 2 Entropy changes of a same osmosis process from the views of different observers. The green arrowhead is the water transferred into the red cell through AQP1. a, from the view of a common observer, it is an entropy decreasing process during which the water is moved against its chemical potential gradient; from the view of AQP1 which can sense the information of molecular size, it is a spontaneous process of entropy increment driven by osmotic pressure. Comparing Equation (2) and (4), it can be found that the entropy functions of a same osmosis process are different for different observers. What's puzzling is that the second law of thermodynamics is broken for an observer while is strictly followed for another. There must be a missing entropy hidden in this puzzling appearance. The molecular size is one kind of information of solute. The special dumbbell shape of AQP1 makes 1'2'()0glucoseureaureadVdsRccdVT it have the ability to "read" this information, which appears as the reflection coefficients of solutes in equation (4). This ability generates an information gap between AQP1 and a common observer without this ability. During the osmosis, this information was continuously consumed. The Landauer's principle linking information and thermodynamics had been confirmed by the experiments of Bérut et al5. By adding the entropy increase caused by the consumption of information, the whole entropy changes from the view of an observer from conventional thermodynamic will also be greater than 0. Taken together, AQP1 in fact can work as one kind of Maxwell's Demon in the body: it can move water against its chemical potential gradient at the cost of consuming information (as shown in Fig.3). The information consumed during the osmosis can be defined as follows, (5) Fig.3. Feedback loop of AQP1 when working as a Maxwell's Demon. dsolute and dpore are the effective diameters of solute molecule and AQP1 vestibule respectively, csolute is solute concentration. Jv is the osmotic water flux driven by the osmotic pressure 1'2'12(1)0ureaureadidsdsRcdV difference(Δπ) and Lp is hydraulic conductivity. The equation describing the relationship between reflection coefficient(σsolute) and relative size of a solute molecule to the pore comes from reference20. In the bottom right corner, there is a diagram for the dumbbell-like shape of AQP1: an extracellular vestibule (blue), a cytoplasmic vestibule (blue), and a narrow channel(pink) connecting them. The measurement of solutes molecular size and concentration by the vestibule of AQP1, the information translation form molecular size to reflection coefficient which (together with solute concentration) determines the osmotic pressure, and the control of osmotic water flux according to osmotic pressure compose a complete feedback loop34 which generates information flow. Comparing with a common thermodynamics observer which can only get the information of solute concentration, AQP1 can work as a Maxwell's Demon in same cases (such as the situation in Fig.2). During the osmosis, the information of the memory of Maxwell's demon is continuously consumed. When σureacurea= cglucos, a thermodynamics balance from the view of AQP1 will be reached. Integrating the above equation from initial state to this balance, we can get the whole information of AQP1 about the molecular size of solutes (I) (6) where V0 is the initial cell volume. 0(1)(1)glucoseureaureaureaureacIRcVc For a normal red blood cell with UT-B, not all information can be used by AQP1 in its osmosis: a part of information (about 23%) loses because of the diffusional urea transport through UT-B during the osmosis. Mutual transformation of information and energy in aquaporin osmosis The information of size effect is a result of the special structure of AQP1, which is controlled by the genetic information in DNA. During the expression of this information, some energy (E1) is consumed. This special structure of AQP1 makes it has the ability to read the information of molecular size of solute. Then, during the osmosis of red blood cells in the kidney with a large gradient of urea, AQP1 can move water against its chemical potential gradient at the cost of consuming the information of size effect: information is converted back into free energy again (E2). In the whole process, information acts as one kind of energy storage medium. (7) Some kinds of information engines, which can convert information into free energy, have been discussed in theory6,8. From the above discussion, an artificial information engine may be realized by a bionics design from the idea of AQP1 osmosis in the kidney. Discussion The way how AQP1 works as a Maxwell's demon confirms the suspicion of James Maxwell about one and a half century ago. However, the special structure of AQP1 is a result of a long period of biological evolution for a better adaptability to the changing 112DNAAQPIEIE environment, knowing nothing about Maxwell's theoretical prediction. Therefore, is there any possible significance of this Maxwell's demon in our body? In the kidney, there is a large urea gradient which is required for the urine concentrating mechanism35. The osmotic water of red blood cells caused by urea can lead to the volume change of red blood cells when they move from the renal cortex to the renal medulla, or vice versa. A smaller reflection coefficient of urea (σurea) for AQP1 means a smaller osmotic water, and then a smaller volume change. This can reduce the eryptosis elicited by osmotic shock. Therefore, a small σurea for AQP1 is necessary to protect the red blood cells from the eryptosis elicited by osmotic shock. In our experiments, for red blood cells of UT-B null mice, the whole volume change caused by urea gradient is only about half of that caused by glucose gradient at a same concentration; for red blood cells of wild mice, the upper limit of this value is 0.61. What's more, the osmotic water of red blood cells caused by urea can also weaken the urea gradient between internal and external the red blood cells and slow down the urea transferring through UT-B. If the osmotically induced water efflux weakens when urea has a small reflection coefficient, this weakening effect of urea transfer can be eased to some extent. Therefore, a small σurea to AQP1 should make a beneficial contribution to the fast urea absorbing and releasing across the erythrocyte plasma membrane through UT-B, which plays an important role in the urine concentrating mechanism36. Materials and Methods Transgenic mice UT-B knockout mice were generated by targeted gene disruption as previously reported37. Wild-type and UT-B null mice at 8~10 weeks old were housed at constant room temperature (23±1°C) and relative humidity (50%) under a regular light/dark schedule (light on from 7:00A.M. to 7:00P.M.) with free accessing food and water before experiments. Reflection coefficients measurement using stopped-flow Fresh red blood cells obtained by tail bleeding (100~200 μl/bleed) were washed three times in phosphate-buffered saline (PBS) to remove serum and the cellular buffy coat. The composition of PBS is (in mM): NaCl, 154; Na2HPO4, 10; NaH2PO4, 10. Stopped- flow measurements were carried out by stopped flow spectrometer SX20 (Applied Photophysics, UK) on a Hi-Tech Sf-51 instrument. For measurement of reflection coefficients, suspensions of red blood cells (~0.5% hematocrit) in phosphate buffered saline were subjected to a 125 mM inwardly directed gradient of urea, mannitol and glucose. The kinetics of decreasing cell volume were measured from the time course of 90° scattered light intensity at 530 nm wavelength 38. Computation of reflection coefficients σs of solutes are calculated by comparing the initial rates of the red blood cell volume change at zero-time39 in different solutions (8) 00(0)fwdVStPvcRTdtV Where V is the cell volume normalized, vw is the partial molar volume of water, S0/V0 is the initial cell surface-to-volume ratio, c is the initial concentration of solute which causes the osmotic challenge23,28,29. For every experiment, the measurement was repeated 10 or 20 times. The zero-time was determined when most scattered light intensities are linear varying and parallel to each other. The dead time was mainly caused by the mixing process of solution and a red blood cell suspension (some new approach has been making to reduce it recently28). A part of signals which deviated significantly from the main tendency were neglected and the average value of the remaining signals was regarded as the result of the experiment. There were three measurements for every solute. For UT-B null red blood cells, a widely used method 25,30–32 developed by van Heeswijk and van Os 1986 40 was also used to calculate reflection coefficients. (9) where y is the scattered light intensity normalized, α, k, c are three parameters. When comparing the water permeability of different channels or membranes to the same osmotic challenge (both the solute and its concentration are the same), k was direct proportional to water permeability (Pf). However, when calculating the σs of different solutes to the same channel (such as AQP1), we needed to use α. (10) ktye000000(1)iooiVbcAVcccc where A is a system parameter, which can be regarded as a constant here;V0 is the initial cell volume, b is the osmotically inactive volume, ci0 is the initial cell osmolarity and co0 is the initial equivalent outside osmolarity, c is the concentration of solute (urea, mannitol or glucose) and σ is its reflection coefficient. The detailed derivation of above equation will be given in support information. Computation of whole volume change Assuming the change of the amount of solutes inside the cell can be ignored when osmotic equation is reached40, we can get (12) (13) From above equation, we can discover that the whole volume change is in direct proportion to α. (14) However, UT-B is an efficient urea channel24. The urea transfer of wild mice red blood cells during the osmosis can't be ignored completely and the whole volume change of red blood cells is reduced to a certain extent. Therefore, for wild mice red blood cells with UT-B, the actual whole volume change is smaller than the value calculated by above equation. 000()()oiVbCVbC00000()(1)iOCVVVbAVC00cos()()ureaureaglucoseglueVVVV Acknowledgements The authors thank A.E Hill (Physiological Laboratory, University of Cambridge), Peter Agre (JHMRI, Johns Hopkins University), and John Mathai (BIDMC, Harvard Medical School) for useful discussions and valuable suggestions through e-mail. The helps in language or illustration from Miss Hu Qiyi, Miss Zhang Yibo and Mr. Oliver Robshaw are also gratefully acknowledged. Reference (1) Maddox, J. Maxwell's Demon Flourishes. Nature 1990, 345, 109–109. (2) Maruyama, K.; Nori, F.; Vedral, V. Colloquium: The Physics of Maxwell's Demon and Information. Rev. Mod. Phys. 2009, 81, 1–23. (3) Serreli, V.; Lee, C.-F.; Kay, E. R.; Leigh, D. A. A Molecular Information Ratchet. Nature 2007, 445, 523–527. (4) Toyabe, S.; Sagawa, T.; Ueda, M.; Muneyuki, E.; Sano, M. Experimental Demonstration of Information-to-Energy Conversion and Validation of the Generalized Jarzynski Equality. Nat. Phys. 2010, 6, 988–992. (5) Bérut, A.; Arakelyan, A.; Petrosyan, A.; Ciliberto, S.; Dillenschneider, R.; Lutz, E. Experimental Verification of Landauer's Principle Linking Information and Thermodynamics. Nature 2012, 483, 187–189. (6) Mandal, D.; Jarzynski, C. Work and Information Processing in a Solvable Model of Maxwell's Demon. Proc. Natl. Acad. Sci. 2012, 109, 11641–11645. (7) Sagawa, T.; Ueda, M. Second Law of Thermodynamics with Discrete Quantum Feedback Control. Phys. Rev. Lett. 2008, 100, 080403. (8) Jayannavar, A. M. Simple Model for Maxwell's-Demon-Type Information Engine. Phys. Rev. E 1996, 53, 2957–2959. (9) Sagawa, T.; Ueda, M. Fluctuation Theorem with Information Exchange: Role of Correlations in Stochastic Thermodynamics. Phys. Rev. Lett. 2012, 109, 180602. (10) Strasberg, P.; Schaller, G.; Brandes, T.; Esposito, M. Thermodynamics of a Physical Model Implementing a Maxwell Demon. Phys. Rev. Lett. 2013, 110, 040601. (11) Mandal, D.; Quan, H. T.; Jarzynski, C. Maxwell's Refrigerator: An Exactly Solvable Model. Phys. Rev. Lett. 2013, 111, 1–5. (12) Agre, P. Aquaporin Water Channels (Nobel Lecture). Angew. Chemie - Int. Ed. 2004, 43, 4278–4290. (13) Law, R. H. P.; Lukoyanova, N.; Voskoboinik, I.; Caradoc-Davies, T. T.; Baran, K.; Dunstone, M. a; D'Angelo, M. E.; Orlova, E. V; Coulibaly, F.; Verschoor, S.; et al. The Structural Basis for Membrane Binding and Pore Formation by Lymphocyte Perforin. Nature 2010, 468, 447–451. (14) Rosado, C. J.; Buckle, A. M.; Law, R. H. P.; Butcher, R. E.; Kan, W.-T.; Bird, C. H.; Ung, K.; Browne, K. a; Baran, K.; Bashtannyk-Puhalovich, T. a; et al. A Common Fold Mediates Vertebrate Defense and Bacterial Attack. Science 2007, 317, 1548–1551. (15) Mueller, M.; Grauschopf, U.; Maier, T.; Glockshuber, R.; Ban, N. The Structure of a Cytolytic Alpha-Helical Toxin Pore Reveals Its Assembly Mechanism. Nature 2009, 459, 726–730. (16) Kiil, F. Kinetic Model of Osmosis through Semipermeable and Solute-Permeable Membranes. Acta Physiol. Scand. 2003, 177, 107–117. (17) Agre, P. Membrane Water Transport and Aquaporins: Looking Back. Biol. Cell 2005, 97, 355–356. (18) Sui, H.; Han, B. G.; Lee, J. K.; Walian, P.; Jap, B. K. Structural Basis of Water-Specific Transport through the AQP1 Water Channel. Nature 2001, 414, 872–878. (19) Gonen, T.; Walz, T. The Structure of Aquaporins. Q. Rev. Biophys. 2006, 39, 361–396. (20) Shu, L.; Liu, X.; Li, Y.; Yang, B.; Huang, S.; Lin, Y.; Jin, S. Modified Kedem-Katchalsky Equations for Osmosis through Nano-Pore. 2014. (21) Rich, G. T. Permeability Studies on Red Cell Membranes of Dog, Cat, and Beef. J. Gen. Physiol. 1967, 50, 2391–2405. (22) Davis, I. S.; Shachar-Hill, B.; Curry, M. R.; Kim, K. S.; Pedley, T. J.; Hill, a. E. Osmosis in Semi-Permeable Pores: An Examination of the Basic Flow Equations Based on an Experimental and Molecular Dynamics Study. 2007. (23) Toon, M. R.; Solomon, A. K. Permeability and Reflection Coefficients of Urea and Small Amides in the Human Red Cell. J. Membr. Biol. 1996, 153, 137–146. (24) Yang, B.; Verkman, a. S. Analysis of Double Knockout Mice Lacking Aquaporin-1 and Urea Transporter UT-B. Evidence for UT-B-Facilitated Water Transport in Erythrocytes. J. Biol. Chem. 2002, 277, 36782–36786. (25) Azouzi, S.; Gueroult, M.; Ripoche, P.; Genetet, S.; Aronovicz, Y. C.; Le Van Kim, C.; Etchebest, C.; Mouro-Chanteloup, I. Energetic and Molecular Water Permeation Mechanisms of the Human Red Blood Cell Urea Transporter B. PLoS One 2013, 8, 2–12. (26) Staverman, A. J. The Theory of Measurement of Osmotic Pressure. Recl. des Trav. Chim. des Pays-Bas 1951, 70, 344–352. (27) Mathai, J. C.; Mori, S.; Smith, B. L.; Preston, G. M.; Mohandas, N.; Collins, M.; van Zijl, P. C. M.; Zeidel, M. L.; Agre, P. Functional Analysis of Aquaporin-1 Deficient Red Cells: THE COLTON-NULL PHENOTYPE. J. Biol. Chem. 1996, 271, 1309–1313. (28) Jin, B.; Esteva-Font, C.; Verkman, A. S. Droplet-Based Microfluidics Platform for Measurement of Rapid Erythrocyte Water Transport. Lab Chip 2015, 15, 3380–3390. (29) Verkman, a. S.; Weyer, P.; Brown, D.; Ausiello, D. a. Functional Water Channels Are Present in Clathrin-Coated Vesicles from Bovine Kidney but Not from Brain. J. Biol. Chem. 1989, 264, 20608–20613. (30) Borgnia, M. J.; Kozono, D.; Calamita, G.; Maloney, P. C.; Agre, P. Functional Reconstitution and Characterization of AqpZ, the E. Coli Water Channel Protein. J. Mol. Biol. 1999, 291, 1169–1179. (31) Kumar, M.; Grzelakowski, M.; Zilles, J.; Clark, M.; Meier, W. Highly Permeable Polymeric Membranes Based on the Incorporation of the Functional Water Channel Protein Aquaporin Z. Proc. Natl. Acad. Sci. U. S. A. 2007, 104, 20719–20724. (32) Yakata, K.; Tani, K.; Fujiyoshi, Y. Water Permeability and Characterization of Aquaporin- 11. J. Struct. Biol. 2011, 174, 315–320. (33) Zhao, D.; Sonawane, N. D.; Levin, M. H.; Yang, B. Comparative Transport Efficiencies of Urea Analogues through Urea Transporter UT-B. Biochim. Biophys. Acta - Biomembr. 2007, 1768, 1815–1821. (34) Strasberg, P.; Schaller, G.; Brandes, T.; Esposito, M. Thermodynamics of a Physical Model Implementing a Maxwell Demon. 2013, 040601, 1–5. (35) Baoxue Yang · Jeff M.Sands. Urea Transporters; Yang, B.; Sands, J. M., Eds.; Subcellular Biochemistry; Springer Netherlands: Dordrecht, 2014; Vol. 73. (36) Bankir, L.; Chen, K.; Yang, B. Lack of UT-B in Vasa Recta and Red Blood Cells Prevents Urea-Induced Improvement of Urinary Concentrating Ability. Am. J. Physiol. Renal Physiol. 2004, 286, F144–F151. (37) Yang, B.; Bankir, L.; Gillespie, A.; Epstein, C. J.; Verkman, a. S. Urea-Selective Concentrating Defect in Transgenic Mice Lacking Urea Transporter UT-B. J. Biol. Chem. 2002, 277, 10633–10637. (38) Yang, B.; Verkman, a S. Urea Transporter UT3 Functions as an Efficient Water Channel. Direct Evidence for a Common Water/urea Pathway. J. Biol. Chem. 1998, 273, 9369–9372. (39) Goldstein, D. a; Solomon, a K. Determination of Equivalent Pore Radius for Human Red Cells by Osmotic Pressure Measurement. J. Gen. Physiol. 1960, 44, 1–17. (40) van Heeswijk, M. P.; van Os, C. H. Osmotic Water Permeabilities of Brush Border and Basolateral Membrane Vesicles from Rat Renal Cortex and Small Intestine. J. Membr. Biol. 1986, 92, 183–193.
1803.06748
1
1803
2018-03-18T21:38:49
Optically resolving the dynamic walking of a plasmonic walker couple
[ "physics.bio-ph" ]
Deterministic placement and dynamic manipulation of individual plasmonic nanoparticles with nanoscale precision feature an important step towards active nanoplasmonic devices with prescribed levels of performance and functionalities at optical frequencies. In this Letter, we demonstrate a plasmonic walker couple system, in which two gold nanorod walkers can independently or simultaneously perform stepwise walking powered by DNA hybridization along the same DNA origami track. We utilize optical spectroscopy to resolve such dynamic walking with nanoscale steps well below the optical diffraction limit. We also show that the number of walkers and the optical response of the system can be correlated. Our studies exemplify the power of plasmonics, when integrated with DNA nanotechnology for realization of advanced artificial nanomachinery with tailored optical functionalities.
physics.bio-ph
physics
Optically resolving the dynamic walking of a plasmonic walker couple Maximilian J. Urban1, Chao Zhou1, Xiaoyang Duan1, and Na Liu1,2* 1Max Planck Institute for Intelligent Systems, Heisenbergstrasse 3, D-70569 Stuttgart, Germany 2Kirchhoff Institute for Physics, University of Heidelberg, Im Neuenheimer Feld 227, D-69120 Heidelberg, Germany ABSTRACT: Deterministic placement and dynamic manipulation of individual plasmonic nanoparticles with nanoscale precision feature an important step towards active nanoplasmonic devices with prescribed levels of performance and functionalities at optical frequencies. In this Letter, we demonstrate a plasmonic walker couple system, in which two gold nanorod walkers can independently or simultaneously perform stepwise walking powered by DNA hybridization along the same DNA origami track. We utilize optical spectroscopy to resolve such dynamic walking with nanoscale steps well below the optical diffraction limit. We also show that the number of walkers and the optical response of the system can be correlated. Our studies exemplify the power of plasmonics, when integrated with DNA nanotechnology for realization of advanced artificial nanomachinery with tailored optical functionalities. KEYWORDS: Artificial nanowalkers, Self-assembly, DNA origami, Dynamic nanoplasmonics, Plasmonic nanostructures, Optical spectroscopy 1 In the cellular world, materials and information throughout the cell are often transported by a myriad of biological motors, which can carry out nanometer steps along protein tracks1. To date, the understanding about how individual motors execute directed movement and how multiple motors work collectively has remained incomplete2-5. Nevertheless, this has not hindered but rather fostered a strong incentive to harness artificial nanomachines for performing functions similar to their biological analogs6. Among diverse opportunities, DNA walkers represent one of the most successful attempts toward this aim7. Over the last decade, miscellaneous DNA walkers, which can perform controlled movement8-19, transport cargo20-21 and direct chemical reactions22, have been accomplished. So far, research efforts on DNA walkers have mostly focused on individual walker behavior. However, implementation of multiple artificial walkers that can walk both individually and collectively on the same track, has not been achieved. By combining recent advances in nanoplasmonics and DNA nanotechnology23-29, in this Letter, we demonstrate a novel dynamic plasmonic walker couple, composed of two gold nanorod (AuNR) walkers. We utilize optical spectroscopy to resolve the dynamic stepwise walking of the walker couple powered by DNA hybridization along a shared DNA origami track. We first show that the two walkers can individually execute nanoscale steps directionally and progressively, demonstrating the independent control of each walker. Subsequently, we show that the two walkers can perform simultaneous walking, working together on the same track. A sensitive plasmonic coupling scheme30 is introduced for in situ optically monitoring the dynamic walking of the two walkers with steps well below the diffraction limit. The two walkers and a stator (an immobilized AuNR) are assembled on the two opposite surfaces of an origami template (see Figure 1a). The two walkers (red and green AuNRs) optically interact with the stator (grey AuNR) simultaneously. In particular, each walker and the stator constitute a chiral 2 geometry31-34. The chiroptical response of the entire system is jointly determined by the positions of both walkers relative to the stator. This renders optical discrimination of the walking directions and the individual steps of the two walkers possible. Finally, to examine the possibility of optically determining the walker number, one of the walkers is dissociated off the track. A clear spectral shift is observed. The experimental results show an overall good agreement with the theoretical calculations. Our study exemplifies the power of plasmonics when integrated with DNA nanotechnology for deterministic placement and active manipulation of plasmonic particles with nanoscale precision, as well as for characterization of dynamic DNA nanostructures using optical spectroscopy. Fig. 1a illustrates the schematic of the plasmonic walker couple system. The nominal dimension of all the AuNRs is 35 nm×10 nm. Double-layer DNA origami35, 36 is utilized to achieve a rigid and robust track. The DNA origami (58 nm×42 nm×7 nm) was prepared following a typical annealing process36 and then purified to remove excess strands (see supporting information). Six rows of footholds (a-f) with equal spacing of 7 nm are extended from the origami template. Both of the walkers are fully covered with foot strands of the same sequences, which are partially complementary to the footholds on the track. In particular, each walker resides on two rows of footholds in order to establish a stable configuration, when reaching a certain system state. The individual walking behavior of the walker couple is first investigated as shown in Fig. 1b. In this case, one walker halts, whereas the other walker carries out stepwise walking. At the initial state, i.e., state I, the red and green walkers are symmetrically positioned at the two ends of the stator. More specifically, the red walker steps on rows a and b, while the green walker steps on rows e and f. The entire system is therefore achiral at state I. When the green walker 3 stays stationary at rows e and f, the red walker can carry out two steps with a step size of 7 nm forward, successively reaching states II and III, following the upper route indicated by the solid arrows (see Fig. 1b). On the other hand, when the red walker stands still at rows a and b, the green walker can perform two discrete steps toward the red walker, reaching system states IV and V, respectively, following the lower route indicated by the dotted arrows (see Fig. 1b). When excited by light, the plasmons generated in the walkers and the stator, are coupled through near-field electromagnetic interactions37-39. Due to the cross configurations formed by the respective walkers and the stator, the collectively coupled plasmons lead to plasmonic circular dichroism (CD). The resulting CD spectra are highly sensitive to the configuration changes of such a three-dimensional (3D) system40. Transmission electron microscopy (TEM) was carried out to access the assembled structures. An overview TEM image of the structures at state I is shown in Fig. 2a, demonstrating a good assembly of the walkers and the stator on the origami. The enlarged TEM image is presented in the black-framed inset of Fig. 2a, where the AuNRs and origami are clearly visible. At state I, the red (green) walker and the stator form a right-handed (left-handed) geometry, therefore rendering the overall state achiral. The CD spectrum at state I was measured using a Jasco-1500 CD spectrometer. The result is shown by a black curve in Fig. 2b. A slight left-handed preference is observable in the CD spectrum. This is likely due to the sample imperfection in the experiment, as any minute assembly deviation may disturb the ideal achiral symmetry, leading to measurable signals resulting from the high sensitivity of CD spectroscopy. The red walker first advances steps toward the green walker as shown by the upper route in Fig. 1b. Such directional walking is activated by toehold-mediated strand displacement processes29,41 as illustrated in Fig. 2c. In brief, the red walker programmably attaches (detaches) 4 its feet on (from) the track through hybridization (de-hybridization) with the corresponding foothold rows. Blocking strands (a') specific to row a, and removal strands (c'') specific to row c are added simultaneously. Dissociation of the red walker's feet from row a is initiated by the blocking strands a' through strand displacement. Meanwhile, the removal strands c'' undergo a strand displacement reaction and take the responsibility to make the footholds on row c accessible for binding the red walker's. As a result, the red walker imposes one step forward in a rolling fashion by stepping on rows b and c, reaching system state II. The corresponding TEM image is shown in the red-framed inset of Fig. 2a. At this state, the asymmetry of the right- handed geometry formed by the red walker and the stator decreases with respect to that at state I, whereas the left-handed geometry formed by the green walker and the stator is unchanged (see Fig. 1b). Therefore, the net handedness of the system is left-handed. This is reflected by a characteristic peak-to-dip profile in the CD spectrum as shown by the red curve in Fig. 2b. Even though it arises from a 7 nm displacement of the red walker on the track, the chiroptical response is very distinct and the CD intensity is as large as 40 mdeg. Subsequently, the red walker executes one more step forward by adding blocking strands b' and removal strands d'', reaching system state III. The red walker and the stator constitute a nominally achiral geometry (see Fig. 1b). The chiroptical response of the system therefore mainly originates from the left-handed geometry formed by the green walker and the stator. As shown in Fig. 2b, the CD intensity increases to approximately 70 mdeg. The TEM images of the structures at station III are presented in the green-framed inset of Fig. 2a. Similarly, the stepwise walking of the green walker can be optically tracked. The red walker resides on rows a and b, while the green walker carries out two successive steps by addition of corresponding blocking and removal strands. The system reaches states IV and V (see the lower 5 route in Fig. 1b). Our system is designed such, that the stepwise walking of the red and green walkers leads to oppositely handed geometries, which are optically distinguishable. More specifically, they are associated with nearly mirrored CD response, when compared the spectra at states II and III with those at states IV and V. In short, through careful structure designs, both the individual walking directions and nanoscale steps of the two walkers on the same track can be optically resolved using standard CD spectroscopy. Theoretical calculations of the CD spectra were carried out using commercial software COMSOL Multiphysics based on a finite element method. The calculated results can be found in supporting information. Overall, the experimental spectra agree well with the theoretical ones. Apart from the individual walking of each walker, our system allows for the simultaneous walking of both walkers. To in-situ monitor the dynamic walking processes as well as to directly compare the simultaneous and individual walking behavior, the timed-dependent CD response was recorded at a fixed wavelength of 695 nm, i.e., at the CD peak or dip position. We examine the individual and simultaneous walking routes, coded A and B, respectively, as illustrated in Fig. 3a. Both routes A and B begin with state I, followed by state II. It is associated with an immediate CD intensity increase as shown by both the dashed (route A) and solid (route B) curves in Fig. 3b. Then, routes A and B start to differ. By following route A (see Fig. 3a), the two walkers carry out two individual steps to reach system state IV. First, the red walker halts, and the green walker executes one step toward the red walker. In fact, it introduces a new state, i.e., state VI. This state is also nominally achiral in that the two walkers are symmetrically positioned at the two sides of the stator. A clear CD decrease is observed as shown by the dashed curve in Fig. 3b. Next, the green walker halts, and the red walker carries out one step backward. The system reaches state IV. This is followed by a further CD decrease as shown by the dashed curve 6 in Fig. 3b. As a result, the transition from state II to state IV via route A is associated with two clear CD decrease steps. Along route B, the two walkers execute simultaneous walking to undergo a direct transition from state II to state IV. The simultaneous walking mechanism is illustrated in Fig. 3c. Blocking strands (c' and f') specific to rows c and f, as well as removal strands (a'' and d'') specific to rows a and d are added at the same time. Through strand displacement processes, the two walkers walk simultaneously by executing a concurrent movement on the track along the same direction. Consequently, the red walker steps on rows a and b, while the green walker steps on rows d and e. The system then reaches state IV. As shown by the solid curve in Fig. 3b, this transition is associated with a single-step, yet large, CD intensity decrease. The overall CD response change via route A is smaller than that via route B, likely due to the influence from the steric hindrance between the two closely spaced walkers at the intermediate state VI. Eventually, both routes A and B end at state V. As shown by the solid and dashed curves in Fig. 3b, the two samples exhibit a similar response with further CD intensity decrease at state V. In short, the walker couple may follow different routes by executing individual or simultaneous walking to reach designated system states. This elucidates a sophisticated control of multiple walkers on the same track, afforded by the high programmability of DNA nanotechnology. Finally, the correlation between the number of walkers and the chiroptical response of the system is investigated. One of the walkers is dissociated off the track. State V is selected as an exemplary starting state because at this state the two walkers are positioned very close to each other. Therefore, the plasmonic coupling aspect between the two walkers can be examined. The TEM images and the CD spectrum (blue curve) at state V are presented in Fig. 4a. By adding the blocking strands (a' and b') specific to rows a and b, the red walker is dissociated off the track 7 (see the left route in Fig. 4b). The green walker still resides on rows c and d. The system becomes nominally achiral. As shown by the green curve in Fig. 4a, this results in a nearly featureless CD spectrum with a slight left-handed preference. In contrast, addition of blocking strands (c' and d') specific to rows c and d dissociates the green walker off the track (see the right route in Fig. 4b). In this case, the red walker remains on rows a and b. As shown by the red curve in Fig. 4a, the resulting CD spectrum exhibits a clear spectral red-shift with a stronger spectral profile, when compared to the blue curve at state V. This is due to the fact that when the red walker is left alone on the track, the transverse coupling between the two walkers vanishes, giving rise to a resonance red-shift according to the plasmon hybridization theory42. The chiroptical response becomes more pronounced because when the green walker is off the track, the left-handed preference disappears and the asymmetry of the system turns larger. In this regard, the number of the walkers on the origami track can be optically resolved. In conclusion, we have demonstrated a plasmonic walker couple system powered by DNA hybridization. We have shown a successful control of two plasmonic walkers, which can independently or simultaneously walk on the same track. The related walking processes can be optically monitored in real time. Also, we have demonstrated that the walker number and the optical response of the system can be correlated. Our studies may outline a very important step toward the development of advanced artificial nanomachinery43,44 with multiple walkers working in concert. From the viewpoint of plasmonics, our studies also offer a solution to deterministic placement and dynamic manipulation of individual nanoparticles with nanoscale precision. This may enable active plasmonic devices with prescribed levels of performance and functionalities, through controlled motion and interaction among multiple plasmonic nanoparticles. FIGURES 8 Figure 1. (a) Schematic of the plasmonic walker couple. The red beam indicates the incident circularly polarized light. (b) Stepwise walking of the individual walkers. At state I, the walkers are symmetrically positioned at the two ends of the stator. Along the upper route, the green walker halts and the red walker takes two successive steps, reaching states II and III, respectively. Along the lower route, the red walker halts and the green walker carry out two successive steps, reaching states IV and V, respectively. 9 Figure 2. (a) Overview TEM image of the plasmonic walker couple structures at state I. Scale bar: 200 nm. Magnified views: structures at states I-V. Scale bar: 30 nm. The plasmonic structures display certain deformation on the TEM grid. (b) Measured CD spectra at different states. (c) Schematic of the walking process from state I to state III. By adding blocking and removal strands, the red walker programmably attaches (detaches) its feet on (from) the track through hybridization (de-hybridization) with the corresponding foothold rows. 10 Figure 3. (a) Routes for individual (I-II-VI-IV-V, route A) and simultaneous (I-II-IV-V, route B) walking. (b) Dynamic processes of individual and simultaneous walking detected by in-situ CD spectroscopy at a fixed wavelength of 695 nm. (c) Schematic of the simultaneous walking process. 11 Figure 4. (a) TEM images and measured CD spectra of the plasmonic walker couple structures at state V and the structures after selective dissociation of one walker off the track. Scale bar: 30 nm. (b) Schematic of the programmable dissociation process. Supporting Information. Materials and methods, theoretical calculations, additional data and detailed walking procedures. This material is available free of charge via the Internet at http://pubs.acs.org. AUTHOR INFORMATION Corresponding Author *E-mail: [email protected] 12 Author Contributions M.U., C.Z., and N.L. designed the project, analyzed the data, and wrote the manuscript. M.U. and C.Z. performed the experiments. X.D. did the theoretical calculations. Notes The authors declare no competing financial interest. Acknowledgements We thank A. Jeltsch and R. Jurkowska for assistance with CD spectroscopy. We thank M. Kelsch and K. Hahn for assistance with TEM microscopy. We thank M. Hentschel and A. Kuzyk for comments on the manuscript. TEM data was collected at the Stuttgart Center for Electron Microscopy (StEM). This project was supported by the Sofja Kovalevskaja grant from the Alexander von Humboldt-Foundation, the Marie Curie CIG grant, and the European Research Council (ERC Dynamic Nano) grant. References 1. Vale, R. D. Cell 2003, 112, 467-480. 2. Endow, S. A.; Higuchi, H. Nature 2000, 406, 913-916. 3. Derr, N. D.; Goodman, B. S.; Jungmann, R.; Leschziner, A. E.; Shih, W. M.; Reck- Peterson, S. L. Science 2012, 338, 662-665. 4. Welte, M. A. Curr. Biol. 2004, 14, R525-R537. 5. Klumpp, S.; Lipowsky, R. Proc. Natl. Acad. Sci. 2005, 102, 17284-17289. 6. 7. 8. 9. 10. Yin, P.; Yan, H.; Daniell, X. G.; Turberfield, A. J.; Reif, J. H. Angew. Chem., Int. Ed. von Delius, M.; Leigh, D. A. Chem. Soc. Rev. 2011, 40, 3656-3676. Pan, J.; Li, F.; Cha, T. G.; Chen, H.; Choi, J. H. Curr. Opin. Biotechnol. 2015, 34, 56-64. Sherman, W. B.; Seeman, N. C. Nano Lett. 2004, 4, 1801-1801. Shin, J. S.; Pierce, N. A. J. Am. Chem. Soc. 2004, 126, 10834-10835. 2004, 43, 4906-4911. 11. Tian, Y.; He, Y.; Chen, Y.; Yin, P.; Mao, C. D. Angew. Chem., Int. Ed. 2005, 44, 4355- 4358. 12. Omabegho, T.; Sha, R.; Seeman, N. C. Science 2009, 324, 67-71. 13. Yin, P.; Choi, H. M. T.; Calvert, C. R.; Pierce, N. A. Nature 2008, 451, 318-322. 14. Lund, K.; Manzo, A. J.; Dabby, N.; Michelotti, N.; Johnson-Buck, A.; Nangreave, J.; Taylor, S.; Pei, R. J.; Stojanovic, M. N.; Walter, N. G.; Winfree, E.; Yan, H. Nature 2010, 465, 206-210. 15. Wang, Z. G.; Elbaz, J.; Willner, I. Nano Lett. 2011, 11, 304-309. 16. Wickham, S. F. J.; Bath, J.; Katsuda, Y.; Endo, M.; Hidaka, K.; Sugiyama, H.; Turberfield, A. J. Nature Nanotech. 2012, 7, 169-173. 17. Green, S. J.; Bath, J.; Turberfield, A. J. Phys. Rev. Lett. 2008, 101, 238101. 18. Loh, I. Y.; Cheng, J.; Tee, S. R.; Efremov, A.; Wang, Z. ACS Nano 2014, 8, 10293- 10304. 19. Tomov, T. E.; Tsukanov, R.; Liber M.; Masoud R.; Plavner N.; and Nir E. J. Am. Chem. Soc. 2013, 135, 11935-11941. 13 20. Gu, H. Z.; Chao, J.; Xiao, S. J.; Seeman, N. C. Nature 2010, 465, 202-205. 21. Cha, T. G.; Pan, J.; Chen, H. R.; Salgado, J.; Li, X.; Mao, C. D.; Choi, J. H. Nature Nanotech. 2014, 9, 39-43. 22. He, Y.; Liu, D. R. Nature Nanotech. 2010, 5, 778-782. 23. Tan, S. J.; Campolongo, M. J.; Luo, D.; Cheng, W. Nature Nanotech. 2011, 6, 268-276. 24. Kuzyk, A.; Schreiber, R.; Fan, Z. Y.; Pardatscher, G.; Roller, E. M.; Hogele, A.; Simmel, F. C.; Govorov, A. O.; Liedl, T. Nature 2012, 483, 311-314. 25. Kuzyk, A.; Schreiber, R.; Zhang, H.; Govorov, A. O.; Liedl, T.; Liu, N. Nature Mater. 2014, 13, 862-866. 26. Schreiber, R.; Luong, N.; Fan, Z.; Kuzyk, A.; Nickels, P. C.; Zhang, T.; Smith, D. M.; Yurke, B.; Kuang, W.; Govorov, A. O.; Liedl, T. Nat. Commun. 2013, 4, 2948. 27. Lan, X.; Chen, Z.; Dai, G.; Lu, X.; Ni, W.; Wang, Q. J. Am. Chem. Soc. 2013, 135, 11441-11444. 28. Tian, Y.; Wang, T.; Liu, W.; Xin, H. L.; Li, H.; Ke, Y.; Shih, W. M.; Gang, O. Nature Nanotech. 2015, 10, 637-644. 29. Zhou, C.; Duan, X.; Liu, N. Nat. Commun. 2015, 6, 8102. 30. Liu, N.; Hentschel, M.; Weiss, T.; Alivisatos, A. P.; Giessen, H. Science 2011, 332, 1407-1410. 31. Hentschel, M.; Schaeferling, M.; Weiss, T.; Liu, N.; Giessen, H. Nano Lett. 2012, 12, 2542-2547. 32. Auguie, B.; Alonso-Gomez, J. L.; Guerrero-Martinez, A.; Liz-Marzan, L. M. J. Phys. Chem. Lett. 2011, 2, 846-851. 33. Yin, X. H.; Schaeferling, M.; Metzger, B.; Giessen, H. Nano Lett. 2013, 13, 6238-6243. 34. Ma, W.; Kuang, H.; Wang, L. B.; Xu, L. G.; Chang, W. S.; Zhang, H. N.; Sun, M. Z.; Zhu, Y. Y.; Zhao, Y.; Liu, L. Q.; Xu, C. L.; Link, S.; Kotov, N. A. Sci. Rep. 2013, 3, 1934. 35. Rothemund, P. W. K. Nature 2006, 440, 297-302. 36. Douglas, S. M.; Dietz, H.; Liedl, T.; Hogberg, B.; Graf, F.; Shih, W. M. Nature 2009, 459, 1154-1154. 37. Halas, N. J.; Lal, S.; Chang, W. S.; Link, S.; Nordlander, P. Chem. Rev. 2011, 111, 3913- 3961. 38. Fan, Z. Y.; Govorov, A. O. Nano Lett. 2010, 10, 2580-2587. 39. Lal, S.; Link, S.; Halas, N. J. Nature Photon. 2007, 1, 641-648. 40. Zhang, S.; Zhou, J. F.; Park, Y. S.; Rho, J.; Singh, R.; Nam, S.; Azad, A. K.; Chen, H. T.; Yin, X. B.; Taylor, A. J.; Zhang, X. Nat. Commun. 2012, 3, 942. 41. Zhang, D. Y.; Seelig, G. Nature Chem. 2011, 3, 103-113. 42. Prodan, E.; Radloff, C.; Halas, N. J.; Nordlander, P. Science 2003, 302, 419-422. 43. Erbas-Cakmak, S.; Leigh, D. A.; McTernan, C. T.; Nussbaumer, A. L. Chem. Rev. 2015, 115, 10081-10206. 44. Browne, W. R.; Feringa, B. L. Nature Nanotech. 2006, 1, 25-35. 14 Table of Contents 15
1805.03432
1
1805
2018-05-09T09:25:43
Crowding and pausing strongly affect dynamics of kinesin-1 motors along microtubules
[ "physics.bio-ph" ]
Molecular motors of the kinesin-1 family move in a directed and processive fashion along microtubules (MTs). It is generally accepted that steric hindrance of motors leads to crowding effects; however, little is known about the specific interactions involved. We employ an agent-based lattice gas model to study the impact of interactions which enhance the detachment of motors from crowded filaments on their collective dynamics. The predictions of our model quantitatively agree with the experimentally observed concentration dependence of key motor characteristics including their run length, dwell time, velocity, and landing rate. From the anomalous stepping statistics of individual motors which exhibit relatively long pauses we infer that kinesin-1 motors sometimes lapse into an inactive state. Hereby, the formation of traffic jams amplifies the impact of single inactive motors and leads to a crowding dependence of the frequencies and durations of the resulting periods of no or slow motion. We interpret these findings and conclude that kinesin-1 spends a significant fraction of its stepping cycle in a weakly bound state in which only one of its heads is bound to the MT.
physics.bio-ph
physics
Crowding and pausing strongly affect dynamics of kinesin-1 motors along microtubules Matthias Rank1 and Erwin Frey1,* 1Arnold-Sommerfeld-Center for Theoretical Physics and Center for NanoScience, Ludwig-Maximilians-Universität München, Theresienstrasse 37, 80333 München, Germany *Correspondence: [email protected] ABSTRACT Molecular motors of the kinesin-1 family move in a directed and processive fashion along micro- tubules (MTs). It is generally accepted that steric hindrance of motors leads to crowding effects; however, little is known about the specific interactions involved. We employ an agent-based lattice gas model to study the impact of interactions which enhance the detachment of motors from crowded filaments on their collective dynamics. The predictions of our model quantitatively agree with the experimentally observed concentration dependence of key motor characteristics including their run length, dwell time, velocity, and landing rate. From the anomalous stepping statistics of individual motors which exhibit relatively long pauses we infer that kinesin-1 motors sometimes lapse into an inactive state. Hereby, the formation of traffic jams amplifies the impact of single inactive motors and leads to a crowding dependence of the frequencies and durations of the resulting periods of no or slow motion. We interpret these findings and conclude that kinesin-1 spends a significant fraction of its stepping cycle in a weakly bound state in which only one of its heads is bound to the MT. INTRODUCTION The collective motion of molecular motors on microtubules (MTs) and their interactions with each other are highly com- plex processes that underlie important intracellular functions. For example, motors of the kinesin-8 family use MTs as molecular tracks along which they perform directed trans- port (1, 2). Having arrived at the MT end, these motors influ- ence the depolymerisation dynamics at this point, and thus have an effect on MT length (2–5) and spindle size (6, 7), properties whose tight regulation is crucial for the normal operation of a cell (8). Kinesin-1 was the first kinesin to be discovered (9), and it is arguably the motor which has been studied in greatest detail. Kinesin-1 is a versatile cargo transporter (10) which uses its two heads (11) to processively walk towards the plus- end of a MT. In the crowded environment of a typical cell, molecular motors and MT-associated proteins (12) compete for a limited number of binding sites on the MTs. As a con- sequence, "traffic jams" consisting of molecular motors may develop on (parts of) the MT (13, 14). A central question is how motors interact with each other in crowded situations like this, and how motors affect each other's ability to bind to and detach from MTs. Several studies have reported (apparently) conflicting results relating to these issues: Thus, Vilfan et al. (15) observed that kinesin motors primarily bind near other motors. Similarly, Muto et al. (16) observed long-range cooperative binding, and Roos et al. (17) discovered that the dwell time of motors increases when they are in the proximity of other motors on the MT. In contrast, Leduc et al. (13) found a reduction in the dwell time of kinesin-8 motors on crowded filaments, in agreement with in vitro measurements of kinesin-1 carried out by Telley et al. (18). How can these findings be reconciled? Firstly, we note that interactions may differ depending on whether motors are mobile (13, 18) or have been immobilized by genetic engineering (15, 17): It appears that an increased dwell time of motors on the MT or cooperative attachment to a MT is primarily found for immobile motors, while mobile motors experience no, or at least less attractive interactions. A second differentiator of these studies was pointed out by Telley et al. (18) who found that the label used to visualise motors by fluorescence microscopy can be crucial. In particular, when these authors failed to reproduce their own earlier results (19) for the crowding behaviour of kinesin-1 using a different label, they concluded that extensive labelling or the use of large labels may lead to non-specific interactions between motors. Therefore, attractive potentials may develop which hold motors on the MT. To minimise these potential effects, Telley et al. removed parts of kinesin's tail (20), such that the motor could still walk with wild-type characteristics (19), and attached a GFP label to only a small proportion of the motors, leaving the vast majority of kinesin motors unlabelled (18). As a consequence, when they varied the abundance of kinesin, they found that this motor's dwell time was inversely related to its (volume) concentration. In our understanding, the situation considered in this study by Telley et al. (18) is closest to the behaviour in an actual cell. Hence, in our theoretical analysis we will mainly compare our results with their data. The Totally Asymmetric Simple Exclusion Process with 1 Langmuir Kinetics (TASEP/LK) (21–23) is commonly em- ployed to describe the collective dynamics of motors on a MT. In this stochastic lattice gas model, motors are described as particles on a one-dimensional lattice (a protofilament of a MT) and step stochastically towards the lattice end. This approach has successfully predicted (22, 23) the existence of traffic jams and domain walls, which were recently observed in experiments (13, 14). Several variations of this stochastic process have considered specific properties of motors, such as their longitudinal (24) or lateral (25) extension. Further- more, additional interactions of motors with each other have been examined (26–28). Among them are so-called mutually interactive Langmuir kinetics (29–31), where binding and unbinding of monomeric particles are directly influenced by the occupation of the nearest-neighbour binding sites. Most of these studies concentrated on fundamental physical prop- erties of the dynamics of motors, such as the different phases of their collective motion. Consequently, the impact of motor- motor interactions on experimentally accessible quantities, such as the motor run length, dwell time, velocity or their numbers of landings (initial attachments) on the lattice per unit length and time, was usually not considered. In this study, we theoretically examine a model which in- cludes motor-motor interactions and a dimeric driven lattice gas. Our aim is to describe the collective motion of processive molecular motors, such as kinesin-1, along a MT. We find that a simple, motor-induced detachment mechanism suffices to quantitatively account for the experimental measurements reported by Telley et al. (18). By developing a mean-field theory, we explore in detail the dependence of motor dwell time, run length, velocity, and landing rate on the volume concentration of kinesin. Furthermore, we find that stochastic pausing of motors on the MT is significantly enhanced by crowding and leads to short-lived traffic jams on the MT, thus recovering the long and frequent periods of interrupted mo- tor motion observed in experiments (18). By comparing the rates of spontaneous detachment and motor-induced detach- ment from the MT, we gain insight into the stepping cycle of kinesin-1, and find that this motor spends a significant fraction (∼ 22%) of its stepping cycle in a weakly bound state. METHODS Monte Carlo simulations We simulate our stochastic lattice gas model with Gillespie's algo- rithm (32), which provides a way of exactly modelling stochastic processes. In the first step, all possible events are collected and statistically weighed with their rates, and an event is randomly cho- sen out of the resulting vector. Another random number is drawn from an exponential distribution with the total rate (i.e., the sum of the rates of all possible events) as the decay parameter, in order to obtain the update time. Subsequently, all rates are updated and the algorithm starts over. In order to account for the long length of Crowding and pausing of kinesin-1 MTs compared to the motors' run length (on the order of 100 steps), periodic boundary conditions were employed on a lattice with 2000 sites. Fitting analytical results to experimental data For the four sets of quantities measured experimentally (18), namely run length, dwell time, velocity, and landing rate of kinesin motors, analytic equations were obtained, see Eqs. (11)–(14). The parame- ters ν (hopping rate of motors) and ωD (their detachment rate) were obtained from the experimental data (18) at low concentrations, as well as the landing rate λ0 of normalised concentration of motors to the MT. In order to obtain the remaining parameters ωA and θ, the analytic results were taken at the concentrations tested in experi- ments, and the deviations from experimental data were weighed by the experimental standard error (18). Subsequently, the sum of the squared weighed errors was taken, and minimized with Mathemat- ica's NMinimize function. In this way, the global fit values ωD and θ are found, see Eq. (18). RESULTS Model description We wish to analyse the stochastic motion of kinesin-1 motor molecules on MTs. Kinesin-1 is a dimer with two heads (11) that can bind to distinct binding sites (33) on two neigh- bouring tubulin dimers (34). Powered by the hydrolysis of ATP (35), it moves processively and unidirectionally (36) to- wards the MT's plus-end (37) along a protofilament (38, 39). It walks hand-over-hand (34), which implies that the rear (lagging) head steps over the front (leading) head to the next binding site in order to complete a step. To describe the collective dynamics of kinesin-1 mo- tors on protofilaments, we employ a one-dimensional lat- tice gas model as illustrated in Fig. 1, where the fluid sur- rounding the MT can be considered as a homogeneous and constant reservoir of motors with concentration c. The corre- sponding mathematical model is based on the Totally Asym- metric Simple Exclusion Process with Langmuir Kinetics (TASEP/LK) (22, 23). Here, we extend it to include the dimeric nature of kinesin-1, and consider an additional in- teraction which accounts for the enhanced detachment of neighbouring motors. To accommodate the extended size of kinesin, and to allow us to adopt simple stepping rules, each motor is described as a rigid particle which simultaneously occupies two sites of a one-dimensional lattice (24). The di- rected motion of motors is modelled as a stepwise stochastic hopping process with rate ν (Poisson process) towards the plus-end (totally asymmetric); stepping is possible only if the target site is not occupied by another motor (exclusion). In the limit of low coverage of a protofilament, each motor would then move at an average speed v0 = νa, where a = 8.4 nm (40) is the size of a tubulin heterodimer. Motors from the reservoir can attach to the protofilament lattice at rate ωA at locations 2 Crowding and pausing of kinesin-1 Figure 1: Lattice gas model for the collective dynamics of kinesin-1 motor proteins moving along a protofilament of a microtubule (MT). Motors are modelled as dimers that simultaneously occupy two neighbouring lattice sites, and advance unidirectionally towards the plus-end (right) of a protofilament at a rate ν (Poisson stepper), if no other motor occupies the next binding site (exclusion process). Kinesin-1 is also assumed to randomly bind to and detach from the protofilament at rates ωA and ωD, respectively. Due to steric exclusion binding is possible only if two adjacent binding sites are empty. In addition to spontaneous detachment with rate ωD, we also account for facilitated detachment of motors that are immediate neighbours. For specificity, we assume that the dissociation rate of the rear motor, i.e., the motor closer to the minus-end (left) is enhanced by a rate θ. where two adjacent lattice sites are empty. This rate depends on the volume concentration of motors as ωA = ωac with a constant ωa. There are two pathways that may lead to the detachment of motors from a protofilament. Firstly, motors may detach spontaneously at a rate ωD. Because this alone cannot explain the decrease in motor dwell time on crowded filaments (18), we secondly assume that motors interact with each other via a process that enhances the detachment rate of motors which are immediate neighbours. Specifically, when two motors meet, we assume that the rear motor's unbinding rate is en- hanced by an additional rate θ; the trailing motor therefore "bounces off" the leading motor, which is consistent with experiments showing that when kinesin runs into an obsta- cle on the MT, the motor (and not the obstacle) is likely to detach (18, 41). The opposite case, where the trailing motor "kicks" the leading motor off the filament, leads to the same phenomena. Alternative scenarios, e.g., enhanced detachment of both motors, have been examined in Ref. (29). Motor currents and density profiles Two central quantities that characterise the collective trans- port of kinesin-1 along MTs are the motor density ρ and the motor current j. In general, both quantities depend on the po- sition along the MT. At the minus-end, the density is expected to show an initial (approximately) linear increase towards a Langmuir plateau due to an "antenna effect" (13, 22, 23): This gradient arises from the combined effects of random motor attachment to and detachment from the MT, as well as driven transport along it; the slope of the initial increase is proportional to the attachment rate ωA. Similarly, a density Figure 2: Bulk motor density and current. Symbols show data obtained from stochastic simulations, the lines depict the results of the mean-field analysis, cf. Eqs. (8) and (10) for parameters ωA = 0.01ν and ωD = ωA/10. (a) The interaction- induced unbinding mechanism reduces the motor density ρ. (b) In contrast, the motor current j reaches a maximum for some finite value of the detachment rate θ. gradient can also be found at the MT's plus-end, in particular for motors which remain bound at this tip for an extended time. Molecular motors with this property include kinesin- 8 (13) and kinesin-4 (14); to the best of our knowledge, no such behaviour has been reported for kinesin-1. Due to (po- tential) gradients at the MT's ends, it is generally difficult to determine the full quantitative behaviour of the motor den- sity (24, 29). One particular property of kinesin-1, the motor in which we are primarily interested in this study, allows for a significant simplification in this respect: its run length (on the order of 1 µm (18)) is significantly less than the length of typical MTs (usually several µm (42)). For this reason, the extent of the gradient region is small relative to the MT length, and the density profile is for the most part spatially uniform on the MT for this motor. By assuming a very long lattice and/or periodic boundary conditions (see Appendix), one can dispense with the specification of the boundary processes. Figures 2a and 2b show the bulk density ρ and current j, respectively, as obtained from stochastic simulations using Gillespie's algorithm (32), see Methods. We find that the ad- ditional detachment of motors facilitated by the interaction between neighbouring motors leads to a monotonic decrease in the bulk density (Fig. 2a) with increasing rate θ; in the limit θ = 0, we recover previous results (24). Interestingly, the motor current shows non-monotonic behaviour as a func- tion of θ (Fig. 2b). There is an optimal value of θ at which the current is maximal. This can be understood in terms of the ability of motor-induced detachment to remove motors from very crowded MTs. Here, the flow of motors is subop- timal due to the emergence of traffic jams, as in the case of vehicular traffic (43). A decrease in the motor density may therefore enhance the numbers of motors transported along the MT per unit time. We will see later that the existence of a maximum motor current follows naturally from the non- monotonic current-density relation, Eq. (10). As an aside, one may thus speculate that motor-induced detachment may serve 3 024/A0.30.4(a)motor density 024/A0.120.140.16(b)motor current j [] to optimise cargo transport along MTs by reducing crowding. In this work, we are mainly interested in examining the collective dynamics of kinesin-1 (9). In experiments, such as those in the study of Telley et al. (18), its collective mo- tion has been characterised in terms of run length on the MT l, dwell time τ, velocity V, and the rate λ (the number of motor landings on the MT per unit time and length). All of these quantities may also be extracted from simulation data. However, not all of the model parameters necessary for sim- ulations can be directly measured in experiments. We will therefore employ the following strategy: First, we develop a theoretical analysis of our model, and extract model parame- ters from experimental data as far as possible. With analytical expressions for all relevant quantities at hand, we then fit our model to the experimental measurements. Eventually, we will show that, with the global fit parameters obtained in this way, the theoretical predictions and simulation data of our model are in excellent agreement with experimental measurements. Mean-field theory The configuration of a lattice at any given instant in time is described by a set of occupation numbers {ni}. A lattice site i (a tubulin heterodimer on the protofilament) is either empty (ni = 0) or occupied by the front head (ni = f ) or back head (ni = b) of a motor dimer. For a statistical description we need the one-site and two-site probabilities, defined as p(i, α) = Prob(ni = α) , p(i, α; j, β) = Prob(ni = α ∧ nj = β) . (1a) (1b) Crowding and pausing of kinesin-1 ni−1 = 0). While an interaction-induced detachment process requires that two dimers are immediate neighbours (ni = f and ni+1 = b), the rate of spontaneous detachment is propor- tional to the single-site probability p(i, f ). In general, the master equation, Eq. (3), is not closed as it links single-site to two-site joint probabilities. However, progress can be made by employing a mean-field approxi- mation that neglects all correlations between the positions of motor dimers other than the steric constraint that dimers are not allowed to overlap, i.e. the front and the back heads of different motors cannot occupy the same lattice site. Further- more, for rigid dimers ni = b implies that site i+1 is occupied by the front head of the same motor, ni+1 = f . In order to show how the two-site joint probabilities can be reduced to one-site probabilities we will consider as an example p(i, f ; i+1, b). This probability, like any joint prob- ability, can be expressed in terms of a conditional probabil- ity: p(i, f ; i+1, b) = p(i+1, bi, f ) p(i, f ). As we are neglect- ing correlations in the position of different dimers, the prob- ability that site i+1 is occupied by the back head of a dimer is independent of whether site i is occupied by the front head of another dimer or empty: p(i+1, bi, f ) = p(i+1, bi, 0). Hence, in a mean-field approximation we have p(i+1, bi, f ) = p(i+1, b(i, f )∨(i, 0)) = p(i+1, b¬(i, b)). Using Bayes' theorem, this can be rewritten in the form p(¬(i, b)i+1, b) × p(i+1, b)/p(¬(i, b)). Here, the remaining conditional prob- ability p(¬(i, b)i+1, b) equals 1 because the states (i, b) and (i+1, b) are mutually exclusive. Hence, we are left with the de- sired decomposition into single-site occupation probabilities: We denote the position of a motor by the position of its front head and define the time-averaged dimer density as p(i, f ; i+1, b) = (2) ρi = p(i, f ) , which is then bounded to ρ ∈ [0, 1 2]. The rate of change of these probabilities can be described in terms of a set of master equations (44). For instance, for the time evolution of the probability that site i is occupied by the front head of a motor, one obtains ∂t p(i, f ) = ν(cid:2)p(i−1, f ; i, 0) − p(i, f ; i+1, 0) + ωA p(i, 0; i−1, 0) − ωD p(i, f ) − θ p(i, f ; i+1, b) . (cid:3) (3) Here, the first term on the right-hand side represents a trans- port current given by the difference between a gain and a loss term. The gain term describes the probability per unit time that a motor (front head of a dimer) located at lattice site i−1 moves forward onto an empty site i, and the loss term describes the probability per unit time that a motor hops from site i to the next (empty) site, i+1. The remaining terms describe attachment and detachment processes with the joint probabilities selecting the allowed lattice configura- tions. Thus, attachment of a dimer to the lattice is possible only if two neighbouring empty sites are available (ni = 0 and p(i+1, b) p(i, f ) = p(i+1, b) p(i, f ) . (4) 1−p(i, b) p(¬(i, b)) Compared to a naive decomposition into single-site occu- pation probabilities p(i+1, b)p(i, f ), this equation includes a factor 1−p(i, b) which corrects for dimers spanning sites i and i + 1, i.e., which takes into account those correlations that are due to the dimeric nature of the motor molecules. In the following we refer to such a factor as the local correlation factor. Using p(i, b) = p(i+1, f ) one may rewrite this result solely in terms of the density ρi as p(i, f ; i+1, b) = (5) . ρi+2 ρi 1−ρi+1 In the same way, cf. Ref. (24), we can also approximate the other joint probabilities of Eq. (3). The ensuing mean-field master equation reads (cid:20) (cid:21) ∂t ρi = ν (1−ρi−ρi+1)ρi−1 + ωA(1−ρi−ρi+1)(1−ρi−1−ρi) 1−ρi 1−ρi − (1−ρi+1−ρi+2)ρi 1−ρi+1 − ωD ρi − θ ρi+2 ρi 1−ρi+1 In the stationary state, where ∂t ρi = 0, this expression recur- sively determines the occupation density of site i in terms of (6) . 4 the densities of the neighbouring sites i±1. In general, the dynamics of such a system is very rich and entails boundary- induced phase transitions (23, 24, 29, 45, 46). As discussed above, kinesin-1 has a relatively short run length and our focus is the bulk of MTs. Hence, we may assume that the motor density is constant, ρi = ρ, and arrive at the mean-field equation ∂t ρ = ωA(1−2ρ)2 ρ2 1−ρ , (7) which yields the motor density ρs in the stationary state (∂t ρ = 0) as 1−ρ − ωD ρ − θ (cid:113) 2ωA 4ωAωD+4θωA+ω2 D ρs = 4ωA+ωD+ . (8) Note that we could also have arrived at Eq. (8) by assuming attachment-detachment balance ωA p(i, 0; i−1, 0) = ωD p(i, f ) + θ p(i, f ; i+1, b) . (9) As we are only interested in the behaviour at steady state, we will omit the index s in the following, i.e. ρ := ρs. By employing the mean-field approximation we can also derive an expression for the motor current j. This quan- tity is defined as the number of motors that pass through a site on the MT per unit time, and is therefore given by ji = νp(i, f ; i+1, 0). By analogy with the derivations of the previous paragraph and Ref. (24), the motor current simplifies to j(ρ) ≈ ν ρ(1−2ρ) 1−ρ . (10) In this equation, we again identify the local correlation factor 1/(1−ρ). Its significance can be understood as follows: Com- pared to the current-density relation for monomeric particles, j(ρ) = ρ(1−ρ), Eq. (10) is skewed, i.e. its maximum lies at a 2(2−√2)≈ 0.29. This density exceeding half-occupation, ρ = 1 agrees remarkably well with the intuitive value for the density 1 3, where on average, every dimer is followed by a vacancy, and is therefore free to jump. With the analytical expressions for the stationary motor density ρ on the MT [Eq. (8)] and their flux j(ρ) [Eq. (10)], we now have a description of the most central physical quan- tities that characterise the collective motion of molecular motors on a MT. As Figs. 2a and 2b show, these analytically calculated quantities agree very well with data from stochastic simulations. Unfortunately, with present-day experimental techniques, it is difficult to measure collective quantities like the density ρ and the current j. It is much easier to determine quantities derived from the observation of single labelled motors. These include the dwell time τ of motors on the MT, their velocity V, run length l, and the landing rate λ. In order to define the link between theory and experiment which we ultimately Crowding and pausing of kinesin-1 aim for, we must therefore also find expressions for these quantities. We first turn to the calculation of the dwell time τ. A motor located at site i can detach either spontaneously at rate ωD, or additionally at a rate θ when another motor is located right next to it at site i+2. The corresponding probability is given by p(i+2, f i, f ), which reduces to ρ/(1−ρ), following the same steps as before. Hence, the dwell time is given by the inverse of the total detachment rate, comprising spontaneous and interaction-induced detachment: (cid:20) (cid:21)−1 τ ≈ ωD + θ ρ 1−ρ . (11) Similarly, in order to obtain the velocity of a motor we need to consider the probability that a particle located at site i finds the next site empty, p(i+1, 0i, f ). This gives for the motor velocity, again using a mean-field approximation, V = V0 p(i+1, 0i, f ) ≈ V0 1−2ρ 1−ρ . (12) With Eqs. (11) and (12), the run length of a motor is given by l = τV ≈ V0 1−2ρ ωD(1−ρ) + θ ρ . (13) Finally, we need to compute the landing rate of kinesin on a MT. In experiments, this quantity is determined by labelling only a small fraction of kinesin, e.g., with GFP, while the vast majority of motors remains unlabelled (18). The con- centration of labelled motors is kept constant at a reference concentration c0, and the unlabelled motors act as crowd- ing agents which are added at varying concentrations. The landing rate is then obtained by counting how many labelled motors land on the MT per unit length and time. In our model, a motor can attach to a site i on the MT only if it finds both site i and the adjacent lattice site i−1 empty, ni = ni+1 = 0. With λ0 being the landing rate of the normalised amount (c0) of labelled kinesin on an otherwise empty MT, the landing rate is λ = λ0 p(i, 0; i−1, 0), which at the mean-field level is approximated by λ ≈ λ0(1 − 2ρ)2 1 − ρ . (14) It is important to note that the normalised landing rate λ0 may differ from ωA(c0). This is because the size of a label such as GFP is comparable to that of the motor. Hence, the attachment rates of labelled and unlabelled motors to the MT may be different. Comparison with experimental data The primary goal of this work is to compare the predictions of our theoretical model with experimental data. Telley et al. (18) have provided an extensive set of measurements for 5 Crowding and pausing of kinesin-1 Figure 3: Comparison with experimental data. Orange circles show the measurements for (a) the run length, (b) dwell time, (c) velocity, and (d) landing rate of kinesin motors, as measured by Telley et al. (18). In blue, we show the fit of our model to this data. Lines are results of our mean field theory, squares compare these calculations with simulations based on Gillespie's algorithm. Figure 4: Characterisation of crowding effects. The plot depicts important physical quantities available from our model, for the same parameters as in Fig. 3. (a) The density of motors on the MT. Because kinesin-1 is a dimer, ρ = 1 2 implies that the lattice is fully decorated with motors. (b) Fraction of detachment events which are due specifically to motor-induced detachment. Even at low concentrations around 7 nM, facilitated dissociation is as prominent as spontaneous detachment. (c) The motor current on the MT, i.e., the number of motors passing over a lattice site per unit time. (d) The landing rate of motors on the MT (orange: experimental data (18), blue: mean-field results), assuming that a single lattice site were sufficient for the landing of a motor. The agreement is worse than for the original model [Fig. 3d]. the motor kinesin-1, which is shown in Fig. 3. Here, the vol- ume concentration of the motor is varied, and this process is incorporated into our model by setting ωA = ωac. From their data, we can directly extract several of our model parameters. The hopping rate ν is obtained from the velocity V0 of a motor in the limit of low motor density (Fig. 3c), ν = 0.66 µm s−1a−1 = 79 s−1 . (15) The detachment rate ωD follows from the dwell time at small motor concentration (Fig. 3b), ωD = 1 1.9 s = 0.53 s−1 , (16) and similarly the landing rate of a normalised amount of labelled kinesin can be directly read off from Fig. 3d at c ≈ 0, (17) This leaves two parameters to be specified, the attachment rate of unlabelled motors to the MT per concentration, ωa, and the rate θ specifying interaction-induced detachment. As λ0 = 1.8 · 10−2 µm−1s−1 . there are four independent sets of quantities that have been measured (18) (run length, dwell time, velocity, and landing rate), comparison of all four with our theoretical results con- stitutes a stringent test of the validity of the assumptions on which the model is based. We have performed a global fit for the four independent quantities l, τ, V, and λ by minimising the squared sum of deviations between experimental measure- ments and mean-field results, weighted by the experimental confidence interval, see Methods. This gives the following values for the rates ωa = 5.4 · 10−2 nM−1s−1 , θ = 2.4 s−1 . (18a) (18b) As can be seen in Fig. 3, using these global fit parameters we find excellent agreement between our theory and all ex- perimentally measured quantities. Both these fit parameters are interesting in themselves. The attachment rate ωa specifies how quickly kinesin attaches to empty lattice sites. In this context, one must keep in mind the fact that the physical quantity underlying the fit is the total motor density ρ on the MT, while the data from Telley et 6 02040motor conc. c [nM]0.40.81.2(a)run length l [m]02040motor conc. c [nM]0.61.01.41.8(b)dwell time [s]02040motor conc. c [nM]0.40.50.60.7(c)velocity V [m/s]02040motor conc. c [nM]0.00.010.02(d)landing rate [m1s1]02040motor conc. c [nM]0.00.10.20.3(a)motor density 02040motor conc. c [nM]04812(c)motor current j[s1]02040motor conc. c [nM]0.40.60.00.2(b)fraction of mot.-ind. det.02040motor conc. c [nM]0.00.010.02(d)landing r. to single site al. (18) are derived from observations of the small minority of labelled motors. In our model, the rate ωa specifies the attach- ment rate of the unlabelled motors, which act as a crowding agent but are otherwise invisible experimentally (18). How then does ωa compare to the landing rate λ0 for labelled mo- tors? This rate was measured at a motor concentration of 5 pM and, assuming that motors in the TIRF setup can walk on roughly half of the 13 protofilaments (41), this can be converted into a per-site attachment rate of approximately 5 · 10−3 nM−1s−1. This value is 10 times smaller than the attachment rate for unlabelled motors, and it demonstrates that, while labelling with GFP conserves many kinetic param- eters of native kinesin (19, 47, 48), the attachment rate of the labelled protein is significantly lower. Secondly, let us look more closely at the rate θ, which quantifies motor-induced detachment from the filament. The value of θ exceeds that of the spontaneous detachment rate ωD by four-fold. This is remarkable, because it implies that, under crowded conditions, motor-induced detachment is the dominant mechanism by which motors leave the MT. We will analyse this and other implications of these parameters in greater detail in the following section. Analysis of crowding effects One strength of our approach to the quantitative description of the collective dynamics of molecular motors with a theo- retical model is that it allows us to infer physical quantities which are experimentally difficult to access. In particular, it is interesting and instructive to study the behaviour of the motor density along the MT, ρ, which is the fundamental quantity characterising the degree of crowding on the MT. In Figure 4a, ρ is plotted as a function of the volume concentra- tion of motors c. At small concentrations, the density rises steeply with c, and becomes half-maximal around 20 nM. At this concentration, on average every second binding site on the MT is occupied by a motor head. As c is increased further, the motor density rises only modestly. This is because attachment of additional motors becomes increasingly un- likely when many motors are already present on the MT, and motor-induced detachment becomes more prominent. Figure 4b shows the fraction of motor detachments in- duced by the presence of another motor, plotted as a function of c. With Eq. (7), we find that the contributions of sponta- neous and motor-induced detachment are already comparable at a motor concentration around 7 nM, significantly below the concentration required for half-occupation [Fig. 4a]. The reason for this is that the rate θ exceeds ωD by several-fold, such that motor-induced detachment plays the central role even on filaments with relatively little crowding. The steep in- crease in the contribution of motor-induced detachment to all dissociation events at low motor concentrations also explains the rapid decrease of quantities such as the motors' run length l [Fig. 3a] and dwell time τ [Fig. 3b] at these concentrations. The motor current j may also be examined directly with Crowding and pausing of kinesin-1 our model and the parameters extracted from experimen- tal measurements (Fig. 4c). Once more, we find a steep in- crease at low concentrations. The current becomes maximal at around c ∼ 20 nM, i.e. the concentration where the den- sity is half-maximal, and for higher concentrations the motor current remains almost constant. Finally, the good agreement of our model with experimen- tal data allows us to study the impact of model variations. For example, it has been suggested (49) that kinesin-1 first binds via a single head to the MT on landing, and subsequently attaches its other head. We have directly tested how a differ- ent attachment mechanism might affect the landing rate by assuming that a single binding site is sufficient for the motor to attach to the MT. As a result, the attachment term in Eq. (7) reduces to ωA(1 − ρ). Fig. 4d compares the landing rate ob- tained in this way with experimental data. Clearly, neither with the fit parameters for the original model, nor with param- eters fitted to the modified model do we obtain satisfactory agreement between theoretical results and experimental data. Therefore, our data suggest that kinesin can land on the MT only where two adjacent binding sites are empty. Crowding alone does not lead to periods of no or slow motion of motors As shown in the previous sections, our mathematical model explains the kinetic data for the run length, dwell time, ve- locity, and landing rate of kinesin-1 motors on MTs with high accuracy. These quantities are averaged over a large number of motors and characterise their collective transport along MTs very well. However, with our model, as well as in experiments, quantities other than averages are also ac- cessible, such as the statistics of individual steps of motors. Such quantities are instructive, as they afford insight into the stochastic motion of kinesin at a deeper level. A particularly interesting finding made by Telley et al. (18) in this respect was that kinesin-1 motors, which normally move at a speeds as high as 79 steps/s along the MT under uncrowded condi- tions, sometimes show periods in which they rest on the MT or their motion is at least considerably slowed down. These periods lasted for several tenths of a second, during which a motor would typically proceed by dozens of steps. It was found that the frequency of these periods increased with the volume concentration of kinesin, and hence with the degree of crowding on the MT (18). However, the authors of that study were only able to im- age the motors every 0.1 s, such that the localisation accuracy of kinesin-1 was of the same order of magnitude as the typical distance traversed between two measurements. Furthermore, because kinesin's stepping mechanism includes chemical re- actions as well as diffusive motion, this motor is a stochastic stepper. Consequently, Telley et al. (18) were faced with the problem of robustly distinguishing periods of no (or very 7 Crowding and pausing of kinesin-1 Figure 5: Periods in which kinesin-1 motors show no or only very slow motion. The model parameters are the same as in Fig. 3. Experimental measurements from Telley et al. (18) are shown in orange. For the detection of these periods, the protocol of Telley et al. (18) was used (see the main text). (a)–(b) Our model cannot explain the frequent periods of no or slow motion of motors observed experimentally. (a) Distance travelled by a motor between the beginning and end of such a period. A threshold value of dc = 5 sites (red line) is too large for reliable detection of these periods: Motors traverse for almost 30 lattice sites between the beginning and end of such a period, which is three times the experimental result (orange line). This implies that most of the detected events actually reflect stochastically slow motion which is otherwise normal, and hence the scheme detects these events inaccurately with this choice of dc. Reduction of the threshold to dc = 3 sites (green), or dc = 2 sites (blue) leads to results that are in closer agreement with experimental data. However, this correspondence deteriorates on addition of Gaussian noise (σ = 20nm) to the simulation data before applying the detection protocol (dashed lines) (b) Duration of the so detected periods of no or slow motion for c = 20 nM and a detection threshold dc = 2 sites. In contrast to the experimental findings, where an exponential distribution was observed, the duration peaks around 0.4 s. This result does not change qualitatively when dc is varied, or Gaussian noise is added at various strengths. (c)–(d) Qualitative agreement with experiments is found when motors can spontaneously switch between an active and inactive mode at rates extracted from experimental data (18). (c) The duration of periods of no or slow motion detected from simulations of this model variant (red squares) is similar to those measured in experiments (18). The duration was computed by extrapolating the (now) approximately exponential distribution of the detected periods below the cutoff time 0.3 s (18) and the main text for details. (d) The per-step probability that a motor is found in a period of no or slow motion. The direct yield from the detection algorithm (red asterisks) is below experimentally observed (18) values. When their frequency is corrected for the time cutoff (red squares), similar to the procedure used in experiments (18), good qualitative agreement is found, in particular at low motor concentrations. Violet triangles show the frequencies obtained with a different algorithm which counts motors that are inactive, or caught up in a traffic jam behind an inactive motor, but not motors which move slowly due to their stochastic motion. The good agreement between these results and the original detection protocol (red squares) reveals that spontaneously inactivated motors are the dominant contribution for periods of no or slow motion. Solid and dashed lines show heuristic estimates of the probability of entering such a period, assuming that motors in a traffic jam require two, or only one binding site on the lattice. For details, see the main text. slow) motion,1 in which motors are assumed to hardly move at all, from stochastically slow motion which simply reflects the stochasticity of kinesin's steps but is otherwise normal. To overcome these problems, Telley et al. (18) developed a detection scheme for the periods of no or slow motion as follows: The location of the motors was measured every 0.1 s. If a motor failed to advance a critical distance dc between two time frames, or its motion was directed off-axis or back- wards (exceeding a critical angle αc), this displacement was considered as a candidate for the onset of a period of no or 1 Note that Telley et al. (18) use the term "pause" for periods in which no or little motion was detected, and they further distinguish between "wait" and "stop" for such events in which kinesin continued its run subsequent to the pause, or detached from the MT. In this work, we distinguish between the phenomenon observed in experiments, which we will call "periods of no or slow motion", and the cause of these periods, which we term "pause" in the following. slow motion. However, in order to mark the start of such a pe- riod, three successive small displacements were required. To account for the effect of experimental noise, single advances exceeding dc were allowed during a period of no or slow motion, so that the period was only considered as terminated when the displacement was greater than dc twice in a row. The key parameters which determine the sensitivity of the detection of periods of no or slow motion are αc and (in particular) dc. On the one hand, these quantities should be chosen to be so large that fluctuations due to experimental noise are unlikely to prematurely terminate such periods. On the other hand, the critical distance must be kept so small that these periods can be robustly distinguished from normal motion which is slow because of the stochasticity of kinesin's steps. With the parameters dc = 40 nm and αc = 60◦, Telley et al. (18) found that approximately every second kinesin mo- 8 MatthiasRankandErwinFreyFigure5:Periodsinwhichkinesin-1motorsshownooronlyveryslowmotion.ThemodelparametersarethesameasinFig.3.ExperimentalmeasurementsfromTelleyetal.(18)areshowninorange.Forthedetectionoftheseperiods,theprotocolofTelleyetal.(18)wasused(seethemaintext).(a)–(b)Ourmodelcannotexplainthefrequentperiodsofnoorslowmotionofmotorsobservedexperimentally.(a)Distancetravelledbyamotorbetweenthebeginningandendofsuchaperiod.Athresholdvalueofdc=5sites(redline)istoolargeforreliabledetectionoftheseperiods:Motorstraverseforalmost30latticesitesbetweenthebeginningandendofsuchaperiod,whichisthreetimestheexperimentalresult(orangeline).Thisimpliesthatmostofthedetectedeventsactuallyreflectstochasticallyslowmotionwhichisotherwisenormal,andhencetheschemedetectstheseeventsinaccuratelywiththischoiceofdc.Reductionofthethresholdtodc=3sites(green),ordc=2sites(blue)leadstoresultsthatareincloseragreementwithexperimentaldata.However,thiscorrespondencedeterioratesonadditionofGaussiannoise(=20nm)tothesimulationdatabeforeapplyingthedetectionprotocol(dashedlines)(b)Durationofthesodetectedperiodsofnoorslowmotionforc=20nMandadetectionthresholddc=2sites.Incontrasttotheexperimentalfindings,whereanexponentialdistributionwasobserved,thedurationpeaksaround0.4s.Thisresultdoesnotchangequalitativelywhendcisvaried,orGaussiannoiseisaddedatvariousstrengths.(c)–(d)Qualitativeagreementwithexperimentsisfoundwhenmotorscanspontaneouslyswitchbetweenanactiveandinactivemodeatratesextractedfromexperimentaldata(18).(c)Thedurationofperiodsofnoorslowmotiondetectedfromsimulationsofthismodelvariant(redsquares)issimilartothosemeasuredinexperiments(18).Thedurationwascomputedbyextrapolatingthe(now)approximatelyexponentialdistributionofthedetectedperiodsbelowthecutotime0.3s(18)andthemaintextfordetails.(d)Theper-stepprobabilitythatamotorisfoundinaperiodofnoorslowmotion.Thedirectyieldfromthedetectionalgorithm(redasterisks)isbelowexperimentallyobserved(18)values.Whentheirfrequencyiscorrectedforthetimecuto(redsquares),similartotheprocedureusedinexperiments(18),goodqualitativeagreementisfound,inparticularatlowmotorconcentrations.Violettrianglesshowthefrequenciesobtainedwithadierentalgorithmwhichcountsmotorsthatareinactive,orcaughtupinatracjambehindaninactivemotor,butnotmotorswhichmoveslowlyduetotheirstochasticmotion.Thegoodagreementbetweentheseresultsandtheoriginaldetectionprotocol(redsquares)revealsthatspontaneouslyinactivatedmotorsarethedominantcontributionforperiodsofnoorslowmotion.Solidanddashedlinesshowheuristicestimatesoftheprobabilityofenteringsuchaperiod,assumingthatmotorsinatracjamrequiretwo,oronlyonebindingsiteonthelattice.Fordetails,seethemaintext.Toovercometheseproblems,Telleyetal.(18)developedadetectionschemefortheperiodsofnoorslowmotionasfollows:Thelocationofthemotorswasmeasuredevery0.1s.Ifamotorfailedtoadvanceacriticaldistancedcbetweentwotimeframes,oritsmotionwasdirectedo-axisorbackwards(exceedingacriticalangle↵c),thisdisplacementwasconsideredasacandidatefortheonsetofaperiodofnoorslowmotion.However,inordertomarkthestartofsuchaperiod,threesuccessivesmalldisplacementswererequired.Toaccountfortheeectofexperimentalnoise,singleadvancesexceedingdcwereallowedduringaperiodofnoorslowmotion,sothattheperiodwasonlyconsideredasterminatedwhenthedisplacementwasgreaterthandctwiceinarow.Thekeyparameterswhichdeterminethesensitivityofthedetectionofperiodsofnoorslowmotionare↵cand(inparticular)dc.Ontheonehand,thesequantitiesshouldbechosentobesolargethatfluctuationsduetoexperimentalnoiseareunlikelytoprematurelyterminatesuchperiods.Ontheotherhand,thecriticaldistancemustbekeptsosmallthattheseperiodscanberobustlydistinguishedfromnormalmotionwhichisslowbecauseofthestochasticityofkinesin'ssteps.Withtheparametersdc=40nmand↵c=60,Telleyetal.(18)foundthatapproximatelyeverysecondkinesinmotorshowedaperiodofnoorslowmotionatsomepointwhileitprogressedalongtheMT.Duringtheperiodsofnoorslowmotion,motorsproceededonaverage10latticesites.Thisvalueseemslarge,butitismuchlessthantheexpected⇠30latticesiteswhichamotorwouldtraverseunderuncrowdedconditionduringtheminimaltimenecessaryfordetectionoftheseperiods(0.3s).InordertocomparethepredictionsofourmodelwiththeexperimentaldataofTelleyetal.(18),weadaptedandappliedtheirexperimentaldetectionschemeforperiodsofno8 tor showed a period of no or slow motion at some point while it progressed along the MT. During the periods of no or slow motion, motors proceeded on average 10 lattice sites. This value seems large, but it is much less than the expected ∼ 30 lattice sites which a motor would traverse under uncrowded condition during the minimal time necessary for detection of these periods (0.3s). In order to compare the predictions of our model with the experimental data of Telley et al. (18), we adapted and applied their experimental detection scheme for periods of no or slow motion to our system. Note, however, that the motion of motors is restricted to a single dimension in our model, while occasional side-steps, as well as off-axis fluctuations are possible in experiments. Consequently, the two parame- ters, dc and αc, used for the experimental detection have to be reduced to a single parameter dc for our purposes. Moreover, since a finite progression dc between two frames was allowed primarily in order to account for experimental inaccuracies which are absent in simulations, dc has to be critically eval- uated, and the role of noise must be simulated. To this end, we first chose the same threshold distance dc = 40 nm as in Ref. (18), corresponding to 5 lattice sites. With this value, we found that the progression of a motor between the beginning and the end of a so defined period of no or slow motion was almost 30 lattice sites [Fig. 5a]. This is significantly larger than the experimentally measured length of 10 lattice sites, and therefore indicates that most of the detected events in fact do not show behaviour which is physically different from normal motion. Thus, most of the periods of no or slow mo- tion detected with this choice of dc result from the stochastic motion of kinesin. Even for a threshold distance of 3 lattice sites, the progression exceeded experimental data, so that we had to reduce the value of dc to 2 lattice sites in order to find agreement with experimental results [Fig. 5a]. However, the agreement found with this parameter choice deteriorated when Gaussian noise was added to the simulation data (in or- der to account for experimental fluctuations) before applying the protocol [σ = 20 nm in Fig. 5a]. Moreover, the statistics of the durations of periods of no or slow motion detected from our simulation data differed from experimental results. While Telley and coworkers (18) report an exponential distribution, our results indicate a non- exponential distribution with peaks around 0.4–0.5 s, see Fig. 5b. Also the addition of Gaussian noise, or variation of the detection threshold dc did not qualitatively change this distribution. We therefore conclude that the detection protocol of Tel- ley et al. (18) is inappropriate for the analysis of the data obtained from stochastic simulations of our original model for two reasons. Firstly, it fails to distinguish reliably between periods of no or slow motion and stochastically slow, but normal motion of kinesin, as the progression of motors be- tween the beginning and end of the detected periods clearly exceeds experimental results. Secondly, the distribution of the durations of periods of no or slow motion in simulations Crowding and pausing of kinesin-1 differs fundamentally from the experimental findings of Tel- ley et al. (18). Consequently, in order to understand the full dynamic behaviour of motors on the MT additional stochastic processes must be taken into account, which are not captured by our original model. This will be the focus of the next section. Spontaneous pausing of motors leads to crowding-dependent frequencies of periods of no or slow motion In order to examine model variations which could possibly explain the experimental findings of Telley et al. (18) on peri- ods in which the motors did not or only very slowly move, we looked at the data they obtained at low motor concentrations. Interestingly, even though motors proceed along the MT (al- most) in the absence of other motors at these concentrations, periods of no or slow motion were observed occasionally. This prompted us to study a variant of our model in which motors can stochastically pause on the MT, i.e., they may temporarily switch to an inactive state in which they can- not move. From the experimental data at low concentrations, we read off a per-step chance of lapsing into inactivity of pinactivation = 0.4%, and a pausing time with average dura- tion T = 0.12 s, after which motors are reactivated again. We therefore introduced rates rinactivation = 0.004ν = 0.32 s−1 and ractivation = 1/0.12 s−1 = 8.3 s−1 at which motors switch to an inactive or active state, respectively. At the molecular level, a motor might become inactive, for instance, when a motor is trapped in an unfavourable chemical state due to imperfect synchronisation of its heads (49); however, the par- ticular molecular mechanism involved is not important for the argument below. If motors are allowed to switch into an inactive mode, we expect crowding to enhance the measured probability of undergoing a period of no or slow motion, because other mo- tors will tend to form a traffic jam behind inactive motors. Although the motors caught up in the traffic jam are not intrin- sically inactive, they are unable to progress until the inactive motor has become active again. Therefore, crowding should amplify the impact of stochastic pausing and consequently lead to frequent periods in which kinesin motors show no or only slow motion along the MT. We tested these expectations directly by performing Monte Carlo simulations of this variant model. Since the two addi- tional stochastic processes, namely spontaneous inactivation and activation of motors, are rare events, we found that they have only a small impact on motor run lengths, dwell time, velocity and landing rate (data not shown). In contrast, motor behaviour changed considerably at the level of individual steps: Unlike the case in our original model, Fig. 1, the dura- tions of periods of no or slow motion were (approximately) exponentially distributed in the variant model, in accordance 9 with experimental findings (18)2. Following Telley et al. (18), it is essential to extrapolate this exponential distribution below the cutoff time 0.3 s in order to obtain the corrected frequency and mean duration of the periods of no or slow motion. The reason for this is that the cutoff 0.3 s is a technical choice, but there is no physical reason why motors would not also experience periods of no or slow motion which are shorter than that. As a result, periods of no or slow motion comprise the detected periods (those lasting 0.3 s and longer), as well as the undetected periods (those of shorter duration). The mean duration of the periods of no or slow motion is there- fore given by the parameter of the exponential decay of the distribution. Figure 5c shows the concentration dependence of the mean duration of periods of no or slow motion, as they were extracted from simulation data in this way, and they re- produce the experimental findings (18) well. Moreover, these values were almost independent of the parameter dc used for the detection algorithm, which ensures that periods of no or slow motion of our model variant are now detected robustly and accurately. We are now in a position to compare simulation data for the frequencies of periods of no or slow motion with those of the experiments of Telley et al. (18), as shown in Fig. 5d. While the uncorrected probabilities (asterisks) remain below experimental values, as expected, the frequencies corrected for the cutoff (squares) are comparable to those found experi- mentally (18) for low concentrations. However, as the concen- tration is increased, we found that the frequencies measured in our simulations exceed experimental values. This points to the need for further modifications of our model. In principle, any additional interactions can be included into our model and data obtained from stochastic simulations. However, a more instructive approach for our purposes is, however, to analyse the physical principles leading to periods of no or slow motion, and explore how exactly the inactivation of a single motor results in the formation of traffic jams which amplify the effect of pausing. To study this, we employed a different algorithm to detect periods of no or slow motion: Here, we only counted motors that were (i) inactive them- selves, or (ii) trapped in a traffic jam behind an inactive motor. In contrast, events in which motors moved slowly because they were caught up in a stochastically assembled traffic jam (in which no motor is inactive) were not taken into account. The frequencies of periods of no or slow motion obtained with this alternative algorithm (triangles in Fig. 5d) agree well 2 As reactivation from an inactive state is a one-step process, the distri- bution of the duration of periods of no or slow motion should be exactly exponentially distributed in the absence of crowding and noise; this agrees with simulation data analysed with the detection algorithm of Telley et al. (18). As the degree of crowding increased due to additional motors on the MT, and as noise was added to the simulation data, the distribution gradually changed and was non-exponential for high crowding and noise level, albeit with an exponential tail for durations > 0.5 s. In order to comply with the procedure of Telley et al. (18), we used the distribution's tail to fit an expo- nential function to the simulation data, as we extrapolated the distribution below the cutoff value 0.3 s in order to obtain, e.g., the mean duration. Crowding and pausing of kinesin-1 with those calculated with the original algorithm (squares in the same Figure). This implies that although stochastically arising traffic jams (in which no motor is inactive) slow down the collective motion of motors (13, 23), they do not increase the incidence of periods of no or slow motion. In contrast, these periods are predominantly due to the spontaneous (and transient) inactivation of motors and the associated formation of traffic jams behind these motors. Given that the dominant cause of periods of no or slow motion is the formation of traffic jams behind inactive motors, further insight can be gained by estimating theoretically the length of these traffic jams. Imagine that a motor pauses at some lattice site. Then, the n-th motor behind this inactive motor is on average n/ρ sites away from it. Since each motor requires two binding sites on the MT, the n-th motor therefore typically has to travel n/ρ − 2n sites to reach the end of the traffic jam. Hence, the time needed for the n-th motor to reach the end of the traffic jam may be estimated as t(n) = (n/ρ−2n)/V. As a consequence, during the time T required for reactivation of an inactive motor, a traffic jam containing N1 = n(T) = TV/(ρ−1−2) motors will form. After the inactive motor has resumed its run, all the motors stuck in the traffic jam can start moving again one after another, so that it will typically take a time N1ν−1 before the original traffic jam has completely dissolved. During this time, another N2 = n(N1ν−1) motors will have reached the end of the traffic jam, and more time will be needed until this additional traffic jam is dispersed, and so on. Taking the sum over the number of motors caught in traffic jams found in this way, the number of motors N = N1 + N2 + . . . which are ultimately affected by a single spontaneously pausing motor is consequently obtained from a geometric series, yielding T v (19) N = ρ−1 − 2 − V ν−1 . This equation suggests that the effect of spontaneous paus- ing is considerably amplified by crowding. While the cause of traffic jams is the inactivation of a single motor, the phe- nomenon detected with the scheme of Telley et al. (18) is also visible for N other motors that are effectively caught in a traffic jam; consequently, pper. no/slow mot. = pinactivation(1+N). Figure 5d shows the probability per step obtained in this way. Given the level of the heuristic arguments, the agreement with simulation data is satisfactory. Having a theoretical estimate for the density dependence, and with Eq. (8) also the concentration dependence, of the frequencies of periods of no or slow motion at hand, further model variations can now be tested in a relatively simple way. For example, it seems plausible that motors align in a traffic jam very compactly, such that each motor requires a single lattice site on the MT only. This would be in accordance with studies in which the decoration of MT sheets with immo- bilised dimeric kinesin was investigated, and it was found that kinesin binds to the MT via a single head only under certain conditions (15, 50). For this model, the n-th motor behind an 10 inactive motor would then have to travel further compared to the original (i.e., spaced) jamming model, namely n/ρ − n sites. In consequence, the term ρ−1−2 in Eq. (19) would be modified to ρ−1−1, and the amplification of spontaneous paus- ing changes accordingly. As shown in dashed lines in Fig. 5d, the resulting per step probability of entering a period of no or slow motion reproduces the experimental concentration de- pendence (18) better than the original model in which motors align spaciously in a traffic jam. In conclusion, we have shown that spontaneous and tran- sient inactivation of motors is the key to an understanding of the occurrence of periods of no or slow motion. The fre- quency of these periods is determined by the formation of traffic jams, in which motors (which are not intrinsically in- active themselves) cannot, or only slowly progress. However, we are at present unable to uniquely determine the precise mechanisms of jamming, and predict quantitatively how ex- actly they amplify the frequencies of periods of no or slow motion of molecular motors. A central problem seems to be that periods of no or slow motion are relatively short-lived compared to the threshold time required to detect such an event. This implies that large numbers of these events remain undetected, and can only be resolved by extrapolating the duration distribution, as explained above. As a consequence, we expect that the estimates of the frequencies of periods in which kinesin motors move only very slowly or come to a complete halt on the MT are subject to relatively large errors. It will in the future therefore be important to further investi- gate the origin of these periods; in particular, algorithms have to be developed which allow a more direct detection of short pauses, e.g., by increasing the frame rate of experiments. Fur- thermore, direct visualisation of the inactive state would be highly informative. In summary, crowding is most probably not the underlying reason for periods of no or slow motion of motors, but acts as an amplifier to increase their frequency, although their ultimate cause is related to inactive states of kinesin motors. The step cycle of kinesin has (at least) two slow transitions Our findings concerning the motor-induced detachment of kinesin motors provide insight into their stepping cycle. We would like to emphasise first that none of the results presented in the previous sections depends on whether disengagement of the front or rear motor from the MT is enhanced by the presence of another motor. Consequently, "bouncing off" (the rear motor detaches) and "kicking off" (the front motor de- taches) interactions lead to identical results (data not shown). In fact, there are experimental indications that it is the trailing motor which bounces off when it encounters another motor on the MT. This was suggested by, among others, Telley et al. (18), who used non-motile rigour mutants, in addition to wild-type kinesin-1. Here, the tightly bound mutant motors act as obstacles on the MT, and the wild-type motors detach Crowding and pausing of kinesin-1 at an enhanced rate on encountering such an obstacle. This would also suggest that when two wild-type kinesin motors come into contact on the MT, it is the trailing motor that is more likely to detach. At the molecular level, these indications enables us to as- sociate the motor-induced unbinding process with a specific state in the mechanochemical cycle of kinesin. This cycle comprises transitions between several states in which one or both kinesin heads are bound to the MT, and the two heads contain different bound nucleotides. During the stepping cy- cle, kinesin passes through a state in which only a single head is bound to the MT. This weakly bound state is reached after the back (i.e., the tethered) head is released from the MT, and the head that remains bound to the MT binds and hydroly- ses ATP. It is likely that this one-head-bound (1HB) state, in which the head attached to the MT is associated either with ADP or ADP·Pi, is the state from which motors usually detach into the cytosol at finishing their run (51, 52). If the lifetime of this state is increased, kinesin should therefore also unbind at an enhanced probability. We hypothesise that the increase in the detachment rate seen when two motors occupy directly adjacent binding sites on the MT is directly related to this weakly bound state. More specifically, when the rear motor's tethered head attempts to step to the next binding site, but finds this site occupied by another motor, the rear motor can leave its 1HB state only by stepping back (which is rare (53)), or by waiting until the next site is vacated. In this case, the back motor is "trapped" in a weakly bound state, and the detachment rate is enhanced accordingly. We, therefore, interpret θ as the dissociation rate of kinesin from the 1HB ADP(·Pi) state. This interpretation is also supported by measurements of the dissociation rate of single-headed kinesin motors which are artificially held in the ADP and ADP·Pi state, where rates of 3.7 s−1 and 3.8 s−1, were found, respectively (51); these measurements are remarkably similar to the value of θ obtained from Eq. (18b). Following these arguments, the time fraction f which a motor spends in the 1HB state during a normal step, may be determined from ωD = f θ. By direct comparison, we obtain f = 0.22, which implies that kinesin-1 remains in the 1HB ADP(·Pi) state for approximately 22% of the time needed to complete a stepping cycle. In summary, our findings suggest that the kinesin-1 step cycle comprises (at least) two transitions which are of similar duration, as opposed to a single rate-limiting step. This is in agreement with a recent interpretation of the kinesin step cy- cle (54). We believe that our study will also help to reconcile conflicting results on the number and type of rate-limiting steps obtained from optical trapping experiments (53, 55), dark-field (56) and interferometric scattering (49) microscopy experiments, as well as from measurements of the statistics of single motor runs (57). While the methods employed in most of these experiments give rise to much shorter length and time scales, labelling of the heads of motors, or applying force to them using an optical trap risks interfering with the 11 step cycle. The advantage of our analysis is that interference effects are minimised. Therefore, crowding experiments (18) provide unique insight into a microscopic process by in a minimally invasive way. DISCUSSION AND CONCLUSION In this work, we have theoretically studied the impact of inter- actions between kinesin-1 motors on their motility and trans- port properties along microtubules. Based on experimental observations, we have generalised a lattice gas model (22, 23) that has previously proven successful in explaining collective phenomena, such as the existence of traffic jams, which have recently been observed experimentally for kinesin-8 (13), and kinesin-4 (14). The generalised model includes the additional process of motor-induced detachment from the microtubule when one motor is directly adjacent to another, as well as the stochastic inactivation (pausing) of motors. With only two fit parameters, namely the rate of motor-induced detach- ment θ, and the attachment rate of motors to empty lattice sites ωA, our model can account for four independent sets of measurements from in vitro experiments (18) with kinesin-1 (Fig. 3). The level of agreement of our model with experimental data allows us to explore the origin of the relatively long pe- riods during which motors hardly move along the MT at all, which have been observed in experiments (18). We find that crowding alone cannot explain the high frequency of these periods ( Fig. 5). We therefore hypothesize that motors may stochastically switch into an inactive mode. Consequently, crowding leads to the formation of traffic jams behind in- active motors; these traffic jams significantly amplify the number of motors which pause on the filament, Eq. (19). Our findings suggest that motors might actually be aligned very densely in a traffic jam (Fig. 5) such that every motor occu- pies only a single tubulin dimer, in accordance with Ref. (15). By comparing the rates of motor-induced detachment and spontaneous unbinding, we find that kinesin-1 motors spend approximately 22% of their stepping cycle in a weakly bound state. Most probably, motor-induced detachment occurs when the rear motor is held in this state for a prolonged time when two motors are directly adjacent, and that its unbinding is therefore increasingly likely. Our approach to quantitatively model the dynamics of molecular motors enables us to investigate collective proper- ties of kinesin-1 motors in a "real life" situation. Firstly, in the experiments of Telley and coworkers (18), on which our model is based, only a small fraction of motors was labelled. Secondly, insight into the interactions of motors with each other has been gained in our study without perturbing motor behaviour by applying forces etc. Our results enable us, for example, to compare the landing rates of labelled and unla- belled motors, and we have found that in fact labelled motors attach to the MT more slowly than unlabelled motors. This illustrates that the choice of a large label can have a crucial Crowding and pausing of kinesin-1 impact on certain quantities, and thus great care should be taken in interpreting experimental data. Most importantly, our model and the experiments of Telley et al. (18) provide unique insight into the stepping cycle of kinesin, which allows us to estimate the lifetime of a specific, weakly bound state. The major drawback of our method is at once its greatest strength: Our approach is very indirect. The application of forces to kinesin motors, e.g. by using optical traps (52, 55), as well as the attachment of large labels such as gold particles to kinesin heads (49, 56) might have crucial influence on motor dynamics (58). Therefore, indirect methods (57, 59) such as the approach employed in this work are essential to confirm, and improve experimental results found by direct observation. Future studies, both theoretical and experimental, will have to examine more closely the formation and dissolution of traffic jams induced by the spontaneous inactivity of a motor, for example. In the same way, the spatial arrangement and conformation of motors in a traffic jam requires closer attention. Such studies are essential to further improve our understanding of the role of interaction between molecular motors for the dynamics along cytoskeletal filaments. This might have important implications for the biological function of such processes in the crowded environments within cells. ACKNOWLEDGEMENTS We thank Ivo Telley and Thomas Surrey for insightful discussion and sharing the experimental data of Telley et al. (18). We also acknowledge support by the DFG via project B02 within SFB 863, as well as by the German Excellence Initiative via the program "NanoSystems Initiative Munich" (NIM). REFERENCES 1. Gupta, M. L., P. Carvalho, D. M. Roof, and D. Pellman, 2006. Plus end-specific depolymerase activity of Kip3, a kinesin-8 protein, explains its role in positioning the yeast mitotic spindle. Nature Cell Biology 8:913–923. 2. Varga, V., J. Helenius, K. Tanaka, A. A. Hyman, T. U. Tanaka, and J. Howard, 2006. Yeast kinesin-8 depolymerizes micro- tubules in a length-dependent manner. Nature Cell Biology 8:957–962. 3. Varga, V., C. Leduc, V. Bormuth, S. Diez, and J. Howard, 2009. Kinesin-8 Motors Act Cooperatively to Mediate Length- Dependent Microtubule Depolymerization. Cell 138:1174– 1183. 4. Melbinger, A., L. Reese, and E. Frey, 2012. Microtubule Length Regulation by Molecular Motors. Physical Review Letters 108:258104. 5. Rank, M., A. Mitra, L. Reese, S. Diez, and E. Frey, 2018. Lim- ited Resources Induce Bistability in Microtubule Length Regu- lation. Physical Review Letters 120:148101. 6. Goshima, G., R. Wollman, N. Stuurman, J. M. Scholey, and R. D. Vale, 2005. Length Control of the Metaphase Spindle. Current Biology 15:1979–1988. 12 Crowding and pausing of kinesin-1 7. Stumpff, J., G. von Dassow, M. Wagenbach, C. Asbury, and L. Wordeman, 2008. The Kinesin-8 Motor Kif18A Suppresses Kinetochore Movements to Control Mitotic Chromosome Align- ment. Developmental Cell 14:252–262. 8. Jordan, M. A., and L. Wilson, 2004. Microtubules as a target for anticancer drugs. Nature Reviews Cancer 4:253–265. 22. Parmeggiani, A., T. Franosch, and E. Frey, 2003. Phase Coexis- tence in Driven One-Dimensional Transport. Physical Review Letters 90:086601. 23. Parmeggiani, A., T. Franosch, and E. Frey, 2004. Totally asym- metric simple exclusion process with Langmuir kinetics. Physi- cal Review E 70:046101. 9. Vale, R. D., T. S. Reese, and M. P. Sheetz, 1985. Identifica- tion of a novel force-generating protein, kinesin, involved in microtubule-based motility. Cell 42:39–50. 24. Pierobon, P., E. Frey, and T. Franosch, 2006. Driven lattice gas of dimers coupled to a bulk reservoir. Physical Review E 74:031920. 10. Hirokawa, N., and R. Takemura, 2005. Molecular motors and mechanisms of directional transport in neurons. Nature Reviews Neuroscience 6:201–214. 25. Melbinger, A., T. Reichenbach, T. Franosch, and E. Frey, 2011. Driven transport on parallel lanes with particle exclusion and obstruction. Physical Review E 83:031923. 11. Hirokawa, N., 1998. Kinesin and Dynein Superfamily Proteins and the Mechanism of Organelle Transport. Science 279:519– 526. 12. Akhmanova, A., and M. O. Steinmetz, 2008. Tracking the ends: a dynamic protein network controls the fate of microtubule tips. Nature Reviews Molecular Cell Biology 9:309–322. 13. Leduc, C., K. Padberg-Gehle, V. Varga, D. Helbing, S. Diez, and J. Howard, 2012. Molecular crowding creates traffic jams of kinesin motors on microtubules. Proceedings of the National Academy of Sciences 109:6100–6105. 14. Subramanian, R., S.-C. Ti, L. Tan, S. A. Darst, and T. M. Kapoor, 2013. Marking and Measuring Single Microtubules by PRC1 and Kinesin-4. Cell 154:377–390. 15. Vilfan, A., E. Frey, F. Schwabl, M. Thormählen, Y.-H. Song, and E. Mandelkow, 2001. Dynamics and cooperativity of mi- crotubule decoration by the motor protein kinesin11Edited by W. Baumeister. Journal of Molecular Biology 312:1011–1026. 16. Muto, E., H. Sakai, and K. Kaseda, 2005. Long-range coopera- tive binding of kinesin to a microtubule in the presence of ATP. The Journal of Cell Biology 168:691–696. 17. H Roos, W., O. Campàs, F. Montel, G. Woehlke, J. P. Spatz, P. Bassereau, and G. Cappello, 2008. Dynamic kinesin-1 clus- tering on microtubules due to mutually attractive interactions. Physical Biology 5:046004. 18. Telley, I. A., P. Bieling, and T. Surrey, 2009. Obstacles on the Microtubule Reduce the Processivity of Kinesin-1 in a Mini- mal In Vitro System and in Cell Extract. Biophysical Journal 96:3341–3353. 19. Seitz, A., and T. Surrey, 2006. Processive movement of single kinesins on crowded microtubules visualized using quantum dots. The EMBO Journal 25:267–277. 20. Berliner, E., E. C. Young, K. Anderson, H. K. Mahtani, and J. Gelles, 1995. Failure of a single-headed kinesin to track parallel to microtubule protofilaments. Nature 373:718–721. 21. Lipowsky, R., S. Klumpp, and T. M. Nieuwenhuizen, 2001. Random Walks of Cytoskeletal Motors in Open and Closed Compartments. Physical Review Letters 87:108101. 26. Celis-Garza, D., H. Teimouri, and A. B. Kolomeisky, 2015. Correlations and symmetry of interactions influence collective dynamics of molecular motors. Journal of Statistical Mechanics: Theory and Experiment 2015:P04013. 27. Chandel, S., A. Chaudhuri, and S. Muhuri, 2015. Collective transport of weakly interacting molecular motors with Langmuir kinetics. EPL (Europhysics Letters) 110:18002. 28. Teimouri, H., A. B. Kolomeisky, and K. Mehrabiani, 2015. The- oretical analysis of dynamic processes for interacting molecular motors. Journal of Physics A: Mathematical and Theoretical 48:065001. 29. Vuijk, H. D., R. Rens, M. Vahabi, F. C. MacKintosh, and A. Sharma, 2015. Driven diffusive systems with mutually inter- active Langmuir kinetics. Physical Review E 91:032143. 30. Messelink, J., R. Rens, M. Vahabi, F. C. MacKintosh, and A. Sharma, 2016. On-site residence time in a driven diffu- sive system: Violation and recovery of a mean-field description. Physical Review E 93:012119. 31. Gupta, A. K., 2016. Collective Dynamics on a Two-Lane Asym- metrically Coupled TASEP with Mutually Interactive Langmuir Kinetics. Journal of Statistical Physics 162:1571–1586. 32. Gillespie, D. T., 1977. Exact stochastic simulation of coupled chemical reactions. The Journal of Physical Chemistry 81:2340– 2361. 33. Nogales, E., S. G. Wolf, and K. H. Downing, 1998. Structure of the αβ tubulin dimer by electron crystallography. Nature 391:199–203. 34. Yildiz, A., 2004. Kinesin Walks Hand-Over-Hand. Science 303:676–678. 35. Hua, W., E. C. Young, M. L. Fleming, and J. Gelles, 1997. Coupling of kinesin steps to ATP hydrolysis. Nature 388:390– 393. 36. Vale, R. D., T. Funatsu, D. W. Pierce, L. Romberg, Y. Harada, and T. Yanagida, 1996. Direct observation of single kinesin molecules moving along microtubules. Nature 380:451–453. 37. Howard, J., and A. a. Hyman, 2003. Dynamics and mechanics of the microtubule plus end. Nature 422:753–758. 13 38. Ray, S., 1993. Kinesin follows the microtubule's protofilament axis. The Journal of Cell Biology 121:1083–1093. 39. Howard, J., 1996. The Movement of Kinesin Along Micro- tubules. Annual Review of Physiology 58:703–729. 40. Hyman, A. A., 1995. Structural changes accompanying GTP hydrolysis in microtubules: information from a slowly hydrolyz- able analogue guanylyl-(alpha,beta)- methylene-diphosphonate. The Journal of Cell Biology 128:117–125. 41. Schneider, R., T. Korten, W. J. Walter, and S. Diez, 2015. Kinesin-1 Motors Can Circumvent Permanent Roadblocks by Side-Shifting to Neighboring Protofilaments. Biophysical Jour- nal 108:2249–2257. 42. Jeune-Smith, Y., and H. Hess, 2010. Engineering the length distribution of microtubules polymerized in vitro. Soft Matter 6:1778. 43. Nagel, K., and M. Schreckenberg, 1992. A cellular automaton model for freeway traffic. Journal de Physique I 2:2221–2229. 44. Weber, M. F., and E. Frey, 2017. Master equations and the theory of stochastic path integrals. Reports on Progress in Physics 80:046601. 45. Lakatos, G., and T. Chou, 2003. Totally asymmetric exclusion processes with particles of arbitrary size. Journal of Physics A: Mathematical and General 36:2027–2041. 46. Shaw, L. B., R. K. P. Zia, and K. H. Lee, 2003. Totally asym- metric exclusion process with extended objects: A model for protein synthesis. Physical Review E 68:021910. 47. Block, S. M., C. L. Asbury, J. W. Shaevitz, and M. J. Lang, 2003. Probing the kinesin reaction cycle with a 2D optical force clamp. Proceedings of the National Academy of Sciences 100:2351–2356. 48. Guydosh, N. R., and S. M. Block, 2009. Direct observation of the binding state of the kinesin head to the microtubule. Nature 461:125–128. 49. Mickolajczyk, K. J., N. C. Deffenbaugh, J. Ortega Arroyo, J. Andrecka, P. Kukura, and W. O. Hancock, 2015. Kinetics of nucleotide-dependent structural transitions in the kinesin- 1 hydrolysis cycle. Proceedings of the National Academy of Sciences 112:E7186–E7193. 50. Moyer, M. L., S. P. Gilbert, and K. A. Johnson, 1998. Path- way of ATP Hydrolysis by Monomeric and Dimeric Kinesin. Biochemistry 37:800–813. 51. Hancock, W. O., and J. Howard, 1999. Kinesin's processivity results from mechanical and chemical coordination between the ATP hydrolysis cycles of the two motor domains. Proceedings of the National Academy of Sciences 96:13147–13152. 52. Milic, B., J. O. L. Andreasson, W. O. Hancock, and S. M. Block, 2014. Kinesin processivity is gated by phosphate release. Proceedings of the National Academy of Sciences 111:14136– 14140. Crowding and pausing of kinesin-1 53. Clancy, B. E., W. M. Behnke-Parks, J. O. L. Andreasson, S. S. Rosenfeld, and S. M. Block, 2011. A universal pathway for kinesin stepping. Nature Structural & Molecular Biology 18:1020–1027. 54. Hancock, W. O., 2016. The Kinesin-1 Chemomechanical Cycle: Stepping Toward a Consensus. Biophysical Journal 110:1216– 1225. 55. Andreasson, J. O., B. Milic, G.-Y. Chen, N. R. Guydosh, W. O. Hancock, and S. M. Block, 2015. Examining kinesin processiv- ity within a general gating framework. eLife 4:1–44. 56. Isojima, H., R. Iino, Y. Niitani, H. Noji, and M. Tomishige, 2016. Direct observation of intermediate states during the stepping motion of kinesin-1. Nature Chemical Biology 12:290–297. 57. Verbrugge, S., S. M. van den Wildenberg, and E. J. Peterman, 2009. Novel Ways to Determine Kinesin-1's Run Length and Randomness Using Fluorescence Microscopy. Biophysical Journal 97:2287–2294. 58. Khataee, H., S. Naseri, Y. Zhong, and A. W.-C. Liew, 2018. Unbinding of Kinesin from Microtubule in the Strongly Bound States Enhances under Assisting Forces. Molecular Informatics 37:1700092. 59. Soza´nski, K., F. Ruhnow, A. Wi´sniewska, M. Tabaka, S. Diez, and R. Hołyst, 2015. Small Crowders Slow Down Kinesin- 1 Stepping by Hindering Motor Domain Diffusion. Physical Review Letters 115:218102. 14
1810.06459
2
1810
2018-12-08T17:20:17
Kinetics of Vesicle Growth
[ "physics.bio-ph", "cond-mat.soft", "q-bio.SC" ]
A mechanism is proposed for the growth of vesicles dispersed in a liquid solvent and a size distribution function is obtained for the vesicles, both from the first principles calculations. This distribution function is shown to be positively skewed and evolving in time obeying a Fokker-Planck type equation. The critical size of the spherical vesicles is shown to grow in time with an exponent of 0.25. A constant is suggested that characterizes how easily a vesicle can absorb amphiphiles into its periphery.
physics.bio-ph
physics
Kinetics of Vesicle Growth Bijit Singha ∗ Department of Physics, Carnegie Mellon University, Pittsburgh, PA 15213 Date: December 11, 2018 Abstract A mechanism is proposed for the growth of vesicles dispersed in a liquid solvent and a size distribution function is obtained for the vesicles, both from the first principles calculations. This distribution function is shown to be positively skewed and evolving in time obeying a Fokker-Planck type equation. The critical size of the spherical vesicles is shown to grow in time with an exponent of 0.25. A constant is suggested that characterizes how easily a vesicle can absorb amphiphiles into its periphery. Contents 1 Introduction 2 Thermodynamics of vesicle formation and growth 3 Size distribution function of the vesicles 4 Time-evolution of the size distribution 5 Conclusion 1 Introduction 1 2 4 5 6 Amphiphilic molecules can aggregate in a solvent to form different interesting structures. The structure of our interest is called vesicles, in which the amphiphilic bilayer can form hollow sphere. The medium inside such a vesicle can have the same or different concentration than the solvent outside but, for our purpose, we will consider same concentration of the solvent both inside and outside the vesicle. The natural assemblance of the amphiphilic bi-molecules to form such symmetric structures, due to its simplicity and homogeneity in composition, can work as a potential tool to understand the dynamics at the mesoscopic scale. Consequently, these systems have been a subject of wide theoretical studies. Vesicles were observed for the first time more than fifty years ago [1] and since then, they have been used to improve our understanding of a number of cellular processes [2 -- 5], and also in various industrial and commercial applications [6]. Along with experimental progress, there have been many important theoretical development in understanding the formation and growth of this biological system. Helfrich, in his famous work on the lipid bilyer elasticity [7], derived the curvature contribution to the free energy and worked on ellipsoidal deformation of the spherical vesicular structure by magnetic fields. Morse and Milner, in their work [8], added the finite size contributions to the free energy derived by Helfrich. They showed that the free energy depends on the vesicle size logarithmically. There were several important works before this as well, arguing the existence of the c ln N contribution to the free energy [9, 10], with c being a constant and is less than zero and N being the number of bi-molecules in the vesicle. Morse and Milner found that c is positive with a value of 7/6 (we will show in a later section how the signature of c will be constrained from the drift of the vesicle size distribution). Several models have been proposed in literature on the positively-skewed size distribution of the vesicles in a solvent [8, 11] as well. This work is motivated by a recent observation that the mode of the vesicle size distribution grows in time with a scaling exponent of 1/4 for some vesicles [12]. This was found by assuming that the characteristic scale associated with this process has a power-law dependence on time, l(t) ∼ tξ, and then ξ was fitted to experimental data. In this work, we find out, from first principles calculations that, for an ideal spherical vesicle, the scaling exponent of the exosome size growth is indeed 1/4 for the critical radius. This helps us to ∗[email protected] 1 develop a formalism to understand the equilibrium size distribution of vesicles at a given time and its time- evolution of such a system that contains vesicles in aggregate of different sizes. We obtain a Fokker-Planck (FP) equation that will dictate the dynamics of the vesicle-growth process. FP equation has been previously used in [13] to understand the formation of spherical vesicles but we will focus primarily on vesicle-growth in this work. To characterize how easily an amphiphilic bi-molecule can assimilate into the periphery of a given vesicle, we also propose a constant in this work. This constant will be determined by the detailed chemical and structural properties of any vesicle. We assume in this work that the dissolution or the growth of the vesicles proceeds through the emission or assimilation of a single amphiphile bi-molecule from the periphery of the vesicles. We have also used the principles of detailed balancing which will hold only when the system is in thermal equilibrium states at all time. Both the assumptions are justified because the vesicle-growth, as found from the experiments, is a sufficietly slow process. Moreover, several interesting properties of the system have direct correspondences to the specific features of the thermodynamic functions, such as free energy. This helps us in two ways: (i) we can come up with the constraints that these functions should satisfy by simply analyzing their consequences to a real system and comparing with the observations, (ii) finding the effect of such functions on the size and time-evolution of the system gives us an intuitive understanding of the corresponding system. This also allows to specify the constants characterizing such processes (e.g. coefficiient of assimilation) from the constraints imposed by observations about the vesicle sizes and their time-evolution. The structure of this paper is as follows. In Section 2, we will find out that the chemical potential corre- sponding to the vesicle growth processes is effectively a constant for large size of vesicles. In Section 3, we will assume a size distribution function for the vesicles in the growth phase and will derive a Fokker-Planck equation using principles of detailed balancing. This will give us an FP equation for the time evolution of the vesicle size distribution in general. In Section 4, we will use this equation for the special case of spherical vesicles to derive the scaling exponent 0.25 of time for the growth process. We will discuss our findings in Section 5. 2 Thermodynamics of vesicle formation and growth When a large number of amphiphile bi-molecules are dissolved, vesicles can be formed in the solvent. In this process, entropy of the solute amphiphiles increases along with an increase in their interaction free energy. The former is due to the increase in the count of all possible microstates of the system, while the second effect is attributed to the fact that the amphiphiles like to be surrounded by similar bi-molecules than the solvent molecules. Structure formation by these amphiphiles becomes possible when the interaction energy dominates the energy associated with the increase in entropy. In what follows in this section, we will show that when such vesicles are formed, the chemical potential of the vesicles is the same for the sufficiently large size. Let's imagine that vesicles in aggregate of different sizes {i} are present in the solvent, with i being the number of amphiphile bi-molecules in a given vesicle, and let's say there are Ni number of vesicles in aggregate-i. From the first law of thermodynamics, we can write d¯Q = dU + P dV, (1) where d¯Q is the heat gained by the system, U is the internal energy, P is the pressure and V is the volume. Because the system is not thermally isolated in general, second law gives us dU + P dV ≤ T dS, (2) where S is the entropy of the system. Because we have different sizes of vesicles, the energy density of the system will depend on the relative amounts of the aggregates and thus, we have to add one more term to the above expression to write dU + P dV −Xi µidNi ≤ T dS, (3) Here µi denotes the chemical potential of a vesicle that belongs to the aggregate-i. The change in Gibbs Free Energy dG = dU + pdV − T dS should be less than zero for any spontaneous process (and is zero for 2 Head(Hydrophilic) Bi-molecule Tail(Hydrophobic) Figure 1: 2-dimensional projection of a vesicle(on the left) with amphiphilic bi-layer. If this vesicle has i number of amphiphilic bi-molecules, it will belong to aggregate-i. Each amphiphile(on the right) has a hydrophobic tail and a spherical hydrophilic head of radius a. processes at equilibrium). Combining this with Eq. (3), we get µidNi ≤ 0. Xi Because vesicle growth is a very slow process, we assume that this is in equilibrium: µidNi = 0. Xi (4) (5) Now, if some vesicles of aggregate-i absorb the same number of vesicles of aggregate-j to form vesicles of aggregate-(i + j), then −dNi = −dNj = +dNi+j. Using this condition in the above equation, we get From the above expression, we can write µi + µj = µi+j. µi + µj+δn = µi+j+δn. (6) (7) Taylor-expanding the second term on the left hand side and the term on the right hand side and ignoring the higher order terms O(δn2) for δn ≪ i, j, we get Using Eq. (6) again gives us µi + µj + δn dµj dn = µi+j + δn dµi+j dn µn = constant (8) (9) for sufficiently large n. This is a very important relation and will determine the equilibrium size distribution of the vesicles. This tells us that the large-sized vesicles will be in equibrium and will not like to exchange amphiphiles, but they can grow in size by absorbing bi-molecules from small vesicles. This will lead to a specific scaling exponent in time-dependence of the distribution function, as will be shown in the next section. 3 3 Size distribution function of the vesicles We assume a distribution function for a vesicle of volume V , containing n bi-molecules at time t. This is similar to a probability density function at a given time if normalized to unity: f (n, V, t) = ψ(n, t)δ(V − V (n)), (10) where the delta function is basically implying that the volume of the vesicle is determined by the number of bi-molecules on its surface. The size-dependence and the time evolution of this function will be dictated by thermodynamics. The size of the vesicle changes by aggregation or dissociation of amphiphile molecules and can be expressed by a kinetic equation ∂f (n, t) ∂t = α(+) n−1,nf (n − 1, t) − α(−) n,n−1f (n, t) + α(−) n+1,nf (n + 1, t) − α(+) n,n+1f (n, t), (11) where α(+) i,i+1 is the average number of events per unit time that one bi-molecule is absorbed by the vesicle to increase the total number from i to i + 1, α(−) i,i−1 is the average number of events per unit time that one bi-molecule is released from the vesicle to decrease the total number from i to i − 1. Defining the flux Jn as Jn = α(+) n,n+1f (n, t) − α(−) n+1,nf (n, t), we can write ∂f ∂t = − (Jn − Jn−1) . (12) (13) The absorption and emission coefficients α(+) and α(−) can be determined from the kinetics of the corre- sponding process. For a vesicle of critical radius, aggregation and dissociation cancels out each other exactly and assuming the following equilibrium distribution of the vesicle f (eq)(n) ∼ exp(cid:20)− Rrev(n) kBT (cid:21) , (14) where Rrev(n) is the work of formation of a vesicle consisting of bi-molecules done in a reversible process. For constant temperature and pressure, this is basically the Gibbs free energy: From this point onwards, we will call ∆G as free energy. Now, we use the principle of detailed balancing to write Rrev(n) = ∆G(n). (15) α(+) n,n+1 α(−) n+1,n = f (eq)(n + 1) f (eq)(n) = exp(cid:20)−(cid:18) ∆G(n + 1) − ∆G(n) kBT (cid:19)(cid:21) . (16) Using these will eventually lead us to a Fokker-Planck equation that can describe time-evolution of a vesicle during the nucleation and growth processes: where ∂f (n, t) ∂t = − ∂J ∂n , J(n, t) = −α(+) n,n+1(cid:20) ∂f (n, t) ∂n + f (n, t) kBT ∂∆G(n) ∂n (cid:21) . (17) (18) Similar expressions are used for a lot of other thermodynamic systems, as shown in [14]. The first term in Eq. (18) is the diffusion flow rate and the second term corresponds to thermodynamic flow rate or the drift flow rate. The first term in Eq. (18) describes the size distribution cloud spreading diffusively while the second term denotes the peak of the same cloud drifting towards a different n-value. In the absence of the 4 second term, this becomes a zero-drift equation with constant diffusion and the peak of the size distribution simply spread out but doesn't move. Thus, we can see from Eq. (18) that it is the size-dependence of the free energy which is pushing the peak of the distribution function to the right and thus dictates the critical growth rate of the vesicles. Let's recall briefly a few ideas about stochastic processes such as the one we are dealing with in this work. We can consider a stochastic differential equation of the form where µ is a K-dimensional vector, σ is an K × K matrix and dWt is a M -dimensional vector that denotes the infinitesimal increment in the Wiener process with the property that dn = µ(n, t)dt + σ(n, t)dWt (19) hWti = 0, hWtWt′i = min(t, t′). (20) (21) Any process corresponding to Eq. (19) is called Ito process. Corresponding to this process, there exists a Fokker-Planck equation satisfied by the conditional probability density P (n, tn0, t0) of the random variable Nt: ∂P (n, t) ∂t = − K Xi=1 ∂ ∂ni [µi(n, t)P (n, t)] + K Xi=1 ∂2 ∂ni∂nj [Dij(n, t)P (n, t)] (22) In Eq. (22), µ is called the drift vector and D = 1 T is called diffusion tensor. We have used tensor indices in Eq. (22) to show the dimensionality of each term explicitly. Defining the diffusion current using Fick's law 2 σσ and the drift current diff(n, t) = −Dij ∂P (n, t) J i ∂ni J i drift(n, t) = µi(n)P (n, t), we can write the Fokker-Planck equation, Eq. (22), as ∂P (n, t) ∂t = − ∂ ∂ni (cid:2)J i drift(n, t) + J i diff(n, t)(cid:3) Comparing Eq. (24) and (25) with Eq. (17) and Eq. (18), we get that the drift current and the drift vector Jdrift(n, t) = αn,n+1 kBT ∂∆G ∂n f (n, t) µ(n, t) = αn,n+1 kBT ∂∆G ∂n . (23) (24) (25) (26) (27) Notice that this framework can be applied generally for any vesicle systems. In the next section, we are going to assume spherical vesicles and derive a more specific equation governing the time-evolution of such a system. 4 Time-evolution of the size distribution We can see from Eq. (18) that we need the free energy of the vesicle in order to find out the size distribution and the dynamic scaling of the vesicle. Free energy is different for different structures but we consider here, for simplicity, an aggregate of spherical vesicles. We use the same form of the free energy for a spherical vesicle as obtained in [8]: ∆G(n) = 8π(κ + 1 2 ¯κ) + kBT (const.) − ckBT ln n . (28) 5 In the above expression, κ and ¯κ are the vesicle's bending rigidity and gaussian rigidity, T is the temperature of the system, n is the number of bi-molecules in a given vesicle, kB is the Boltzmann constant and c is another constant. Because the first term in the right hand side is scale-independent, the phase behavior is dictated only by size-dependent contribution coming from the last term in that expression. Notice that, the logarithmic dependence of that term tells us that the chemical potential is effectively a constant near the critical radius of the vesicle, as expected from Eq. (9). Using Eq. (17) and (18), and assuming that the diffusivity remains constant during the growth process (αn,n+1 = D0 for sufficiently large n), we get ∂ψ ∂t = D0 ∂ ∂n (cid:20) ∂ψ ∂n + ψ kBT ∂∆G ∂n (cid:21) . (29) Comparing this expression with Eq. (22), (24), (27) with the above equation, we get the drift speed of the size distribution function assumed in Eq. (10): Here nM denotes the mode size at a given time. Thus, u(n, t) = ∂nM ∂t = cD0 nM . nM (t) = p2cD0 t1/2 ∼ t1/2 . (30) (31) (32) Comparing the above expression with the experimental result, we can get the diffusivity constant corre- sponding to this process. We can express the same differential equation in terms of the radius of the vesicle using n = 4πR2 πa2 = 4R2 a2 , where R is the radius of the vesicle and can be measured experimentally, and a is the radius of the cross- section of the hydrophilic head of the vesicle. This gives the the dependence of the critical radius on time: Rc = a 2 (cid:18) 7D0 3 t(cid:19)1/4 ∼ t1/4 . (33) The scaling exponent of time in the above equation is 0.25. It is worth mentioning here that, in [12], the authors found a similar result by assuming a power-law scaling of the mean diameter and fitting the exponent to their experimental data. Also, our observable is the size of the vesicles and it will be more helpful to express Eq. (29) in terms of the radius of the vesicles. and thus, from Eq. (29), ∂ψ ∂t = 5 Conclusion dn = 8RdR a2 , D0a4 64R2 (cid:20) ∂2ψ ∂R2 −(cid:18) 1 + 2c R (cid:19) ∂ψ ∂R + 4cψ R2 (cid:21) . (34) (35) We have given a framework under which we can calculate the time evolution of size distribution of the vesicles, and showed that the critical size of sphericle vesicles will grow with time with a scaling exponent of 0.25. This is derived in a straightforward way, first by assuming that the free energy is only logarithmically (i.e. slowly) varying with the number of bi-molecules, thus the chemical potential being a constant effectively for sufficiently large size of vesicles, and then finding out how the size distribution will evolve by writing down the Fokker-Planck equation corresponding to the process of bi-molecule exchange between different aggregates. Furthermore, we have proposed a constant to characterize the ease of diffusivity of an amphiphile bi-molecule into the periphery of the vesicle. The chemistry and the detailed structural properties will be encoded in this constant while the growth process will be dictated by thermodynamics. This means any deviation from the ideal value of this constant (i.e. value obtained from this work for a spherical vesicle with same solvent with similar concentration inside and outside) can give us information about the physical or chemical interactions or any other form of perturbations playing a role in the system. All these behooves further investigation with more complicated vesicle systems. The scaling exponent will vary similarly for non-ideal structures but we can use the same underlying theory mentioned in this article to comment on the compositions and geometry of the corresponding vesicles, scope for future work. 6 Acknowledgements The author is thankful to Saigopalakrishna Yerneni and Sushil Lathwal for all the valuable discussions about vesicles and some other interesting biological systems, and to Prof. Changjin Huang and Prof. Markus Deserno for their valuable comments on this work. This work was supported by Carnegie Mellon University Department of Physics. References [1] A. D. Bangham, M. M. Standish, J. C. Watkins. Diffusion of univalent ions across the lamellae of swollen phospholipids. the Journal of Molecular Biology 13 (1965) 238 [2] D. P. Pantazatos and R. C. Macdonald. Directly Observed Membrane Fusion Between Oppositely Charged Phospholipid Bilayers. The Journal of Membrane Biology 170 (1999) 27 [3] A-S. Cans et. al. Artificial cells: Unique insights into exocytosis using liposomes and lipid nanotubes. Proceedings of the National Academy of Sciences 100 2 (2003) 400 [4] G. Lei and R. C. Macdonald. Effects on interactions of oppositely charged phospholipid vesicles of covalent attachment of polyethylene glycol oligomers to their surfaces: adhesion, hemifusion, full fusion and "endocytosis". The Journal of Membrane Biology 97 (2008) 221 [5] M. M. Lapinski, A. Castro-Forero, A. J. Greiner, R. Y. Ofoli and G. J. Blanchard. Comparison of liposomes formed by sonication and extrusion: rotational and translational diffusion of an embedded chromophore. Langmuir 23 (2007) 11677 [6] P. Fernandez, N. Willenbacher, T. Frechen, A. Kuhnle. Vesicles as rheology modifier. Colloids and Surfaces A: Physicochemical and Engineering Aspects 262 (2005) 204 [7] W. Helfrich. Elastic Properties of Lipid Bilayers: Theory and Possible Experiments. Z. Naturjiorsch. C 28 (1973) 693 [8] D. C. Morse and S. T. Milner. Fluctuations and Phase Behavior of Fluid Membrane Vesicles. Europhysics Letters 26 (1994) 565 [9] W. Helfrich. Size distributions of vesicles : the role of the effective rigidity of membranes. Journal of Physics (Paris) 47 (1986) 321 [10] D. Huse and S. Leibler. Phase behaviour of an ensemble of nonintersecting random fluid films. Journal of Physics (Paris) 49 (1988) 605 [11] V. Guida. Thermodynamics and kinetics of vesicles formation processes. Advances in Colloid and Inter- face Science 161 (2010) 77 [12] M. Paulaitis, K. Agarwal, P. Nana-Sinkam. Dynamic Scaling of Exosome Sizes. Langmuir 34 (2018) 9387 [13] E. H. Zapata, L. M. Balbuena, I. S. Holek. Thermodynamics and dynamics of the formation of spherical lipidic vesicles. Journal of Biological Physics 35 (2009) 297 [14] V. V. Slezov. Kinetics of First-order Phase Transitions. Wiley-VCH (2009) 7
1008.0194
2
1008
2010-12-13T22:44:24
Swimmers in thin films: from swarming to hydrodynamic instabilities
[ "physics.bio-ph", "cond-mat.soft" ]
We investigate theoretically the collective dynamics of a suspension of low Reynolds number swimmers that are confined to two dimensions by a thin fluid film. Our model swimmer is characterized by internal degrees of freedom which locally exert active stresses (force dipoles or quadrupoles) on the fluid. We find that hydrodynamic interactions mediated by the film can give rise to spontaneous continuous symmetry breaking (swarming), to states with either polar or nematic homogeneous order. For dipolar swimmers, the stroke averaged dynamics are enough to determine the leading contributions to the collective behaviour. In contrast, for quadrupolar swimmers, our analysis shows that detailed features of the internal dynamics play an important role in determining the bulk behaviour. In the broken symmetry phases, we investigate fluctuations of hydrodynamic variables of the system and find that these destabilize order. Interestingly, this instability is not generic and depends on length-scale.
physics.bio-ph
physics
Swimmers in thin films: from swarming to hydrodynamic instabilities Department of Mathematics, University of Bristol, Clifton, Bristol BS8 1TW, U.K. Marco Leoni and Tanniemola B. Liverpool (Dated: October 17, 2018) We investigate theoretically the collective dynamics of a suspension of low Reynolds number swim- mers that are confined to two dimensions by a thin fluid film. Our model swimmer is characterized by internal degrees of freedom which locally exert active stresses (force dipoles or quadrupoles) on the fluid. We find that hydrodynamic interactions mediated by the film can give rise to spontaneous continuous symmetry breaking (swarming), to states with either polar or nematic homogeneous order. For dipolar swimmers, the stroke averaged dynamics are enough to determine the leading contributions to the collective behaviour. In contrast, for quadrupolar swimmers, our analysis shows that detailed features of the internal dynamics play an important role in determining the bulk be- haviour. In the broken symmetry phases, we investigate fluctuations of hydrodynamic variables of the system and find that these destabilize order. Interestingly, this instability is not generic and depends on length-scale. In nature, baths of micron-scale swimmers are found to show remarkable out of equilibrium phenomena ranging from anomalous diffusion and viscosity enhancement, to turbulent and swirl-like motion or self organization into complex dissipative structures [1 -- 5] . They act as the in- spiration for man-made devices able to control and mix fluids on micron scales. The search for the design prin- ciples of such devices [6 -- 8] remains in its infancy. How- ever, simplified models of low Reynolds number swim- mers should prove to be a useful starting point for the theoretical understanding of this class of collective phe- nomena. A very simple picture [9] then, of a swimmer is an inter- nally driven vector oriented towards its direction of mo- tion. Theoretically the dynamics of a collection of such objects can be described, on long length- and time-scales, by vector and tensor equations which are natural gener- alizations of equilibrium liquid crystalline hydrodynam- ics [10]. These active fluids involve the study of conserved and broken symmetry variables that are non-equilibrium analogues [11] of Goldstone modes. Previous studies on active suspensions [12 -- 15] have shown that fluctua- tions in these modes, destabilize ordered states. In 3d this instability is termed generic, as it is independent of length-scale. However, these descriptions are in fun- damental sense, phenomenological since another physical mechanism [14, 15] must be invoked to generate the or- dered states which have subsequently been shown to be unstable [12]. In this letter, we introduce and study a self-contained and microscopically defined physical system in which it is possible to both generate homogeneously ordered states and to examine their stability. This is provided by a 'suspension' of swimmers confined to two dimensions by a viscous thin film [5]. In addition the thin film geom- etry is particularly accessible to experiments both from the point of view of ease of observation and external acti- vation. We study analytically a model that directly links the collective behaviour to the microscopic dynamics. It combines fluctuations, both active and passive, with the deterministic motion due to activity and hydrodynamic interactions. We find that the purely physical coupling mediated by the thin film can give rise to local spontaneous breaking of symmetry and we identify the possible ordered states. The system can have both nematic order ; characterized by a macroscopic axis of mean orientation n and symme- try n → −n, or polar order ; with mean orientation axis p for which p 6= −p. We then examine the stability of each of the homogeneous phases to hydrodynamic fluctu- ations. We find that the isotropic phase is stable. On the other hand, as for the bulk 3d system, the homogeneous broken symmetry states are destabilized by the hydro- dynamic modes of the system. Here, however the thin film weakens this effect and the instabilities observed are length scale dependent (i.e. not generic). We restrict ourselves to a dilute solution, and we make use of a mean field approximation and study c(R, u, t) = the one-particle distribution function: Pi hδ(Ri − R)δ( ui − u)i, the probability of finding a swimmer with average orientation u at position R. This satisfies a dynamic equation ∂tc = −∇ · JT − R · J R ; JT = −DT · ∇c + (¯v u + V)c, J R = −DRRc + Ωc, (1) in terms of translational JT (R, u, t) and rotational J R(R, u, t) currents, with R := u ∧ ∂ ∂ u [16]. DT and DR are respectively the translational (rotational) diffu- sion tensor (constant) of the swimmer and represent both passive and active fluctuations. The deterministic quan- tities are active, describing self-propulsion: ¯v u, and the translational V and rotational Ω velocities induced on a swimmer due to the activity of others by hydrodynamic interactions mediated by the film. Thin film hydrodynamics: The film is described as an infinite incompressible two dimensional layer of fluid with (2d) viscosity η filling the plane z = 0 and coupled hydro- dynamically to another incompressible bulk fluid of (3d) viscosity ηe which fills the region z 6= 0. To distinguish between them we indicate the three-dimensional quanti- ties with a prime. We consider the fluid dynamics in the vanishing Reynolds number (Stokes) limit where inertia can be neglected [17]. For the in-plane quantities, given an in-plane force density F (x, y) the velocity v(x, y) and pressure p(x, y) satisfy [18] η∇2 ⊥ v + ∇⊥p + σ+ e − σ−e = −F ; ∇⊥ · v = 0 (2) where ∇⊥ = (∂x, ∂y) is the 2d gradient operator. σ±e := ηe∂z v′0± is the shear stress of the bulk fluid at the top/bottom of the thin film; in the external region, z 6= 0, the velocity v′(x, y, z) and pressure p′(x, y, z) satisfy the Stokes equation : ηe∇2v′ + ∇p′ = 0 ; ∇ · v′ = 0 , where ∇ = (∇⊥, ∂z) is a 3d gradient operator. The ratio of the two and three dimensional viscosities introduces a length-scale s := η/(2ηe) that governs two asymptotic regimes. For r ≪ s dissipation occurs almost entirely in plane and hydrodynamic flow fields are quasi-two dimen- sional while at lengths r ≫ s dissipation is mostly due to flow out of plane, and the hydrodynamics is similar (but not identical) to that in three dimensions The viscous drag coefficient on a flat disk of radius a embedded in the film is γ = 4πη/g , with g a function of s/a [18]. On length scales large compared to a, the interaction between several disks lying in the film can be approximated using point-like forces at their centers and the Green's function H of eq. (2), corresponding to the flow, v(r) = H(r − r0) · f0 generated by a point-like force f0 at r0. The tensor H(r) = s is the thin film equivalent of the Oseen tensor [16]. In the following, we work in the limit r ≫ s [19, 20] . (2π)2 e−ik·r (I−k⊗k) η R d2k sk2+k We consider swimmers with an average speed v mov- ing in a direction u which can be represented at large length-scales as a time-dependent force dipole that gener- ates an associated velocity field. We also consider swim- mers for which the force dipole is zero and the leading behaviour is determined by a force quadrupole. Hence the average force density of a swimmer with mean posi- tion, rα is of the form fα(r) = −f dL uui∇iδ(r − rα) + 1 uui uj∇i∇jδ(r − rα) + · · · , representing the dipole 2 f qL and quadrupole contributions respectively where L is the typical dimension of the swimmer. 2 For a concrete calculation, an explicit microscopic model of a swimmer is required and we have used a three-disk model of swimmer [21]. This is character- ized by a minimum number (2) of degrees of freedom, with typical length l that move in a non-reciprocal fash- ion in time with frequency ω = 2π T to achieve loco- motion in the Stokes limit. In the following we indi- cate the time-average over a swimmer cycle period T with an over-bar h = 1 In the limit of small sinusoidal oscillations of amplitude d around l is possible to obtain [21] the average self-propulsion ve- locity as v = a0ωd2 (1+κ)2 ] where a0 := 4s 6l2 κ2 − 1 T R T 0 h(t)dt. [1 + 1 3g 2 κ 12l2 [( 1 36l2 [−1 + 2 κ − 1) + 2( 1 κ + 2κ + 2κ2 + 2 and our convention is that the swimmer move in the same direction of u. For the average force we find κ2 − κ) + (1−κ) L = l and f d = f da0 (1+κ)2 ] κ2 + 2 f q = f da0 (1+κ)2 ]. Here f := γωd is the force scale on each degree of freedom of the swimmer. The parameter κ, which is the average ratio of the internal lengths, controls the nature of the swimmer: the dipole is positive (pusher) for κ > 1, neg- ative (puller) for κ < 1 and zero (quadrupole) for κ = 1. The interaction between two such swimmers, which is the origin of V, Ω in eq. (1), is complex as each one is characterized by periodic internal dynamics whose cy- cles may also have different phases [22]. Here we will restrict ourselves to swimmers that all have the same phase [22], and the limit where the typical separation r between the swimmers centres RA and RB is much larger than the typical dimension l of the object. Dy- namical quantities depending on hydrodynamic interac- tions, such as translational v2b := Rα − v uα and an- gular ω2b := uα ∧ uα velocities felt by each swimmer, for α = A, B, are described to good approximation by the first few terms of the expansion in spherical harmon- ics [15, 22] of the tensor H. To leading order, for swim- mer A we find v2b = sΘ(2) r2 }+O(1/r3) and ω2b = sΘ(2) r3 + (3AB − C) u } + O(1/r4). Here A := (r · uA), B := (r · uB), C := ( uA · uB) where uA and uB are swimmers orienta- tions. r is the separation vector and Θ(2) is related to the time-dependent force dipole and scales as f l. We neglect the effect of interactions on the internal dynamics and so do not address synchronization effects [20]. r2 −B uB 2πη {3(A + 2BC − 5AB2) uA∧ r 2πη {(3B2 −1) r A∧ u r3 B The collective dynamics is obtained by a coarse grain- ing procedure, first in time and then in space. The ve- locities Ω, V in eqs. (1) are obtained from the two body velocities for swimmers with positions R and R′ and ori- entations u and u′ as (cid:18) Ω V (cid:19) (R, u) = Z u′,R′ (cid:18) ω2b v2b (cid:19) (R − R′, u, u′)c(R′, u′) (3) where, as before, the over-bar denotes time average over the swimmer cycle. Order parameters such as local den- sity ρ, polarization P and nematic orientation tensor S are defined as moments of c :   ρ P S   (R, t) =Z u   1 u ( u ⊗ u − I/2)   c(R, u, t). (4) Homogeneous ordered states These are states in which c and its moments defined above do not vary with position (denoted by c0, ρ0, P0, S0 respectively). Under these conditions, the mean field velocities are Ω0 = β0 u ∧ P0 + β1 u ∧ S0 · u and V0 = βT P0 where the coefficients β0, β1 and βT are averaged quantities d 1 1 s η 1 4λ a0d l2 (1+κ)2 [ 1 that depend on the microscopic details of each swim- l (κ2 − 1) and mer [23]. To leading order β0 ≈ f 3 (κ − 1) + 5 β1 ≈ f s κ2 )]. η They are both positive for κ > 1 (pushers) and negative for κ < 1 (pullers). Inserting the expressions of Ω0 and V0 in eq (1) , taking the time derivative of eq (4) , we obtain dynamic equations for the moments of c0. Density is conserved, hence ∂tρ0 = 0. The others are, (16)2λ3 12 (κ2 − 1 6 (κ3 − 1 κ ) + 1 ∂tP0 = − DRP0 + β0 2 ρ0P0 +(cid:18) β1 ∂tS0 ab = − 4DRS0 ab + β0(cid:16)P 0 a P 0 b − − β0(cid:19) S0 · P0; δab(cid:0)P0)2(cid:17) + β1 2 2 1 2 (5) ρ0S0 ab. Analysis of eqs (5) shows that, when β0 or β1 are pos- itive, the system can undergo a bifurcation, that sig- nals the appearance of order. Hence pushers can give rise to order whereas pullers cannot. The conditions −DR + ρ0β0/2 = 0 and −4DR + ρ0β1/2 = 0 define two critical lines in the space of parameters ρ0 and f , above which the instability occurs [24]. The ratio β1/(4β0) quadrupoles + N 40 20 0 ρ I 0.3 I 0.6 f 4 0 ρ P N unstable pullers -0.6 stable I I stable pushers 0 -0.3 0.3 0.6 f FIG. 1: Phase diagram showing homogeneous states and their hydrodynamic stability. Isotropic-Nematic and Nematic- Polar transitions for dipolar swimmers. The transitions can occur only for pushers. Force f is measured in arbitrary units. Parameters are set to s η = DR = 1, d = 3a0, l = 3d, λ = 2, and κ = 1.2. In the inset is shown the I-N transition for quadrupolar swimmers (same parameters except κ = 1) . determines if the I-N or I-P transition occurs at lower density (see Fig. 1 where we have chosen parameters for which I-N occurs first). For pullers β0 and β1 are negative and contribute to enhance the noise (diffusion) in the system. For quadrupolar swimmers, these lead- ing order terms vanish and in addition to higher order terms one must keep track of internal mode dynamics of the swimmers. The leading order terms in β0 are nega- tive, so enhance diffusion while in β1 they are positive and promote order. Hence, for this particular micro- scopic prescription, we find no polar phase. We find β1 ≈ f s l (2π + 1)] [20]. A plot of the η critical line in this case is shown in the inset of Fig. 1. Clearly, this transition occurs at higher densities than dipolar swimmers since it is due to higher order terms. 384λ3 [1 + d a0d2 l3 7 We now discuss fluctuations in hydrodynamic variables about isotropic and ordered states for dipole-swimmers. 3 1 As in many active systems, we find that their effect is to destabilize order [12]. In the following we con- sider the deviations of the fields from their homogeneous values given by δρ = ρ − ρ0, similarly for P and S. We introduce Fourier transforms in the usual way as Isotropic state (2π)2 e−ik·r f (k), f (k) = Rr eik·rf (r). f (r) = Rk In the homogeneous isotropic state ρ = ρ0, P0 = S0 = 0 and ρ is the only hydrodynamic variable. Variables δ Pk := δ P · k, splay δS kk and bend k⊥)a := δSbckc(δab − kakb) show diffusive behaviour. (δS From the resulting set of equations density fluctuations are linearly stable [14]. Polar state In the homogeneous polar state ρ = ρ0, P = P0 and the hydrodynamic variables are ρ and the P . The orientation tensor is slaved to P, and director given by S = SP P 2( P P − I 2 ) where SP is determined by eqn. (5). The magnitude P is not a hydrodynamic variable and relaxes to a constant value on microscopic time-scales. We set P = 1 in the following and study linear perturbations around this state. Decomposing k = cos φ P 0 + sin φk⊥, with k⊥ · P 0 = 0, to leading or- der in k we find that fluctuations in density and director P ), giving a growing mode are coupled via splay (k⊥ · δ Γ with real part ℜ(Γ) ∼ k sSP ¯Θ(2) cos 2φ(2 + cos 2φ) and imaginary part, which determines the propagation speed, ℑ(Γ) ∼ ±k ¯v√2 sin φ, as shown in figure 2(a). When ℜ(Γ) is positive fluctuations grow exponentially, signalling an instability of the ordered state due to hydrodynamic in- teractions, analogous to those found in 3d [12]. However unlike those, the growth rate here scales as k in the k → 0 limit. 8η Nematic state A similar analysis can be performed around the homogeneous nematic state (finite ρ0, S0). In this case hydrodynamic variables are density, ρ and nematic orientation tensor, S = S( n ⊗ n − I 2 ). Again the magnitude S relaxes fast and in the following will be set to S = 1. To leading order in k fluctuations of density δ ρ and nematic director δ n are decoupled and splay fluctu- ations have a real mode Γ ∼ k s ¯Θ(2) 4η cos 2φ(2 + ρ0 cos 2φ). For angles above π/4 destabilize the order, as shown in Fig. 2 (b). For quadrupolar swimmers the relevant terms describing hydrodynamic interactions scale as k2 instead of k. Their analysis is not reported here [20]. It is instructive to compare our results with previous studies of swimmers in a 3d fluid [12, 14, 15]. Our analy- sis starts from a microscopic model, and 'integrates out' the fluid degrees of freedom to see the effect on the other hydrodynamic modes. Alternatively one may perform a phenomenological analysis of an ordered state as in [12] replacing the Stokes equation by eq (2); this yields qualitatively the same hydrodynamic instabilities that we have presented above. In essence, the thin film model changes the Fourier spectrum of the hydrodynamic ker- nel from k2 to k + sk2 which in the limit sk ≪ 1 reduces to ≈ k. This is the origin of the different scaling of the in- ) Γ ( e R 1 0 -1 -2 -3 0 1 0.5 0 -0.5 ) Γ ( m I 1 2 φ (rad) 3 ) Γ ( e R 2 1 0 -1 -2 -3 0 1 2 φ (rad) 3 (b) -1 0 2 4 6 φ (rad) 8 10 (a) FIG. 2: Angular dependence of the growing mode Γ in the splay fluctuations around homogeneous order for pushers. A (a) polar state. ℜ(Γ) is positive sign indicates instability. (2) measured in units of k(sSPΘ )/(8η) and ℑ(Γ) in units of k¯v/√2; (b) nematic state. ℜ(Γ) is measured in units of k(s ¯Θ(2))/(4η) and we have used ρ0 = 0.3. stability here. In another contrast, hydrodynamic inter- actions of simplified far-field models of swimmers cannot lead to homogeneous order (swarming) in 3d [15]. This can be ascribed to a mathematical cancellation that oc- curs performing angular integrals of spherical harmonics generated by the hydrodynamic kernels. In the thin film limit considered here, this is circumvented by the con- finement of the swimmer directors to two dimensions. In addition the instability generated by activity in the film is 'soft' in the sense that it scales with k → 0 in compar- ison to the 3d case where it is independent of k (hard) to leading order [12]. We can conclude that in comparison to 3d, hydrodynamic interactions between swimmers in a thin film favour order but are still not strong enough to overcome the destabilizing effect of activity. Finally we note that our approach is complementary to models where simple phenomenological rules of interaction be- tween swimmers are used to study aspects of collective behaviour [11, 24, 25]. In contrast, here we 'derive' such rules from a particular microscopic model using a coarse- graining procedure which is valid under a precise set of conditions. The question of how to characterise the system be- yond the instability [14, 25] remains open. This high- lights a fundamental theoretical issue. The instabilities of homogeneous ordered states seen here and in other non-equilibrium active systems illustrate the limits of the Landau-Ginzburg framework [12, 15], which has been so successful in the study of phase transitions in equilib- rium systems. It might be that alternative approaches, such as describing the homogeneous state as dynamical rather than stationary, or the development of a formalism in which hydrodynamic and non-hydrodynamic variables are treated on the same level, have to be considered. ML acknowledges the support of University of Bristol research studentship. TBL acknowledges the support of the EPSRC under grant EP/G026440/1. 4 [1] X.-L. Wu and A. Libchaber, Phys. Rev. Lett. 84, 3017 (2000). [2] C. Dombrowski, L. Cisneros, S. Chatkaew, R. E. Gold- stein, and J. O. Kessler, Phys. Rev. Lett. 93, 098103 (2004); I. Tuval, L. Cisneros, C. Dombrowski, C. W. Wol- gemuth, J. O. Kessler, and R. E. Goldstein, PNAS 102, 2277 (2005). [3] I. H. Riedel, K. Kruse, and J. Howard, Science 309, 300 (2005). [4] K. C. Leptos, J. S. Guasto, J. P. Gollub, A. I. Pesci, and R. E. Goldstein, Phys. Rev. Lett. 103, 198103 (2009). [5] A. Sokolov, I. S. Aranson, J. O. Kessler, and R. E. Gold- stein, Phys. Rev. Lett. 98, 158102 (2007). [6] R. Dreyfus, J. Baudry, M. Roper, M. Fermigier, H. Stone, and J. Bibette, Nature 437, 862 (2005). [7] P. Dhar, T. M. Fischer, Y. Wang, T. E. Mallouk, W. F. Paxton, and A. Sen, Nano Letters 6, 66 (2006). [8] A. Ghosh and P. Fischer, Nano Letters 9, 2243 (2009). [9] T. J. Pedley and J. O. Kessler, Annual Review of Fluid Mechanics 24, 313 (1992). [10] P. G. de Gennes and J. Prost, The physics of liquid crys- tals (Oxford University Press, 1995). [11] T. Vicsek, A. Czir´ok, E. Ben-Jacob, I. Cohen, and O. Shochet, Phys. Rev. Lett. 75, 1226 (1995). [12] R. Aditi Simha and S. Ramaswamy, Phys. Rev. Lett. 89, 058101 (2002). [13] I. S. Aranson, A. Sokolov, J. O. Kessler, and R. E. Gold- stein, Phys. Rev. E 75, 040901 (2007). [14] D. Saintillan and M. J. Shelley, Phy. Rev. Lett. 100, 178103 (2008); Phys. Fluids 20, 123304 (2008). [15] A. Baskaran and M. C. Marchetti, PNAS 106, 15567 (2009). [16] M. Doi and S. Edwards, The Theory of Polymer Dynam- ics (Oxford University Press, 1986). [17] E. M. Lifshitz and L. D. Landau, Course of Theoreti- cal Physics, Volume VI: Fluid Mechanics, 2nd Edition (Butterworth-Heinemann, 1987). [18] P. G. Saffman and M. Delbruck, PNAS 72, 3111 (1975); P. G. Saffman J. Fluid. Mech. 73, 593-602 (1976);D.K. Lubensky and R.E. Goldstein, Phys. Fluids 8, 843 (1996). [19] A.J. Levine, T.B. Liverpool and F.C. MacKintosh, Phys. Rev. Lett. 93, 038102 (2004); Phys. Rev. E 69, 021503 (2004). [20] M. Leoni and T. B. Liverpool, unpublished(2010). [21] A. Najafi and R. Golestanian, Phys. Rev. E 69, 062901 (2004); R. Golestanian and A. Adjari, Phys. Rev. E 77, 036308 (2008). [22] G. P. Alexander, C. M. Pooley, and J. M. Yeomans, Jour- nal of Physics: Condensed Matter 21, 204108 (2009). [23] To obtain these we perform integrals which require a regularization at small distances. This cut-off is naturally provided by the swimmer size λl, for λ ≥ 2. ical Journal - Special Topics 157, 111 (2008). [24] F. Peruani, A. Deutsch, and M. Bar, The European Phys- [25] H. Chat´e, F. Ginelli, G. Gr´egoire, and F. Raynaud, Phys. Rev. E 77, 046113 (2008).
1601.05295
1
1601
2016-01-20T15:15:21
A novel computational modelling to describe the anisotropic, remodelling and reorientation behaviour of collagen fibrres in articular cartilage
[ "physics.bio-ph", "q-bio.TO" ]
In articular cartilage the orientation of collagen fibres is not uniform, varying mostly with the depth on the tissue. Besides, the biomechanical response of each layer of the articular cartilage differs from the neighbouring ones, evolving through thickness as a function of the distribution, density and orientation of the collagen fibres. Based on a finite element implementation, a new continuum formulation is proposed to describe the remodelling and reorientation of the collagen fibres under arbitrary mechanical loads: the cartilaginous tissue is modelled based on a hyperelastic formulation, being the ground isotropic matrix described by a neo-Hookean law and the fibrillar anisotropic part modelled by a new anisotropic formulation introduced for the first time in the present work, in which both reorientation and remodelling are taken into account. To characterize the orientation of fibres, a structure tensor is defined to represent the expected distribution and orientation of fibres around a reference direction. The isotropic and anisotropic constitutive parameters were determined by the good validation of the numerical models with the experimental data available from the literature. Considering the effect of realistic collagen fibre reorientation in the cartilage tissue, the remodelling algorithm associated with a distribution of fibres model showed accurate results with few numerical calculations.
physics.bio-ph
physics
manuscript No. (will be inserted by the editor) A novel computational modelling to describe the anisotropic, remodelling and reorientation behaviour of collagen fibres in articular cartilage S. Cortez · A. Completo · J. L. Alves Received: date / Accepted: date Abstract In articular cartilage the orientation of collagen fibres is not uniform varying mostly with the depth of the tissue. Besides, the biomechanical response of each layer of the articular cartilage differs from the neighbouring ones, evolving through thickness as a function of the distribution, density and orientation of the collagen fibres. Based on a finite element implementation, a new continuum formulation is proposed to describe the remodelling and reorientation behaviour of the collagen fibres under arbitrary mechanical loads. The cartilaginous tissue is modelled based on a hyperelastic formulation, being the ground isotropic matrix described by a neo-Hookean law and the fibrillar anisotropic part modelled by a new anisotropic formulation introduced for the first time in the present work, in which both reorientation and remodelling are taken into account. To characterize the orientation of fibres, a structure tensor is defined to represent the expected distribution and orientation of fibres around a reference direction. The isotropic and anisotropic constitutive parameters were determined by the good validation of the numerical models with the experimental data available from the literature. Considering the effect of realistic collagen fibre reorientation in the cartilage tissue, the remodelling algorithm associated with a distribution of fibres model showed accurate results with few numerical calculations. Keywords articular cartilage · anisotropy · collagen fibres remodelling · finite element analysis S. Cortez (B) · J. L. Alves Department of Mechanical Engineering University of Minho Guimaraes, Portugal E-mail: scortez,[email protected] A. Completo Department of Mechanical Engineering University of Aveiro Aveiro, Portugal 6 1 0 2 n a J 0 2 ] h p - o i b . s c i s y h p [ 1 v 5 9 2 5 0 . 1 0 6 1 : v i X r a 2 1 Introduction S. Cortez et al. In tissue engineering, the mechanical properties of the engineered material are crucial for any application and/or treatment. The mechanical stimuli imposed on bioreactors are between the most important factors on the tissue growth. Finite element analysis based on computer models for articular cartilage is commonly used to understand the intrinsic growth factors (collagen fibres orientation, solute transport, growth of extracellular matrix constituents and others) and to simulate the effect of changes of the tissue structure under well-defined loading conditions (Chung and Ho, 2010; Nava et al, 2013; Hossain et al, 2014; Bandeiras et al, 2015). In order to improve the mechanical properties of the tissue-engineered cartilage, the organization of the collagen network associated with the mechanical loading needs a deeper attention. The collagen fibres in the articular cartilage have a non-uniform orientation and distribution varying with the depth of the tissue. Because of this inhomogeneous structure, each cartilage layer responds differently to the same mechanical load (Pierce et al, 2013; Federico and Herzog, 2008; Khoshgoftar et al, 2013; Gasser et al, 2006; Federico and Gasser, 2010). The role of fibres is essentially mechan- ical, promoting tissue stiffness and strength. Besides, the mechanical behaviour of connective tissues is strongly influenced by the structural arrangement of the fibres. Consistent with classical arcade-like descriptions of collagen network organi- zation in cartilage (Wilson et al, 2004), fibres in the superficial zone are preferably oriented in the plane parallel to the articular surface, while in the middle zone fibres are dispersed with a random orientation. In the deep tissue zone, fibres tend to be oriented perpendicularly to the bone cartilage interface, i.e., perpendicular to the articular surface (Pearle et al, 2005; Wilson et al, 2007). Several nonlinear constitutive models have been proposed to study the mechan- ical behaviour of tissues containing collagen fibres and some of constitutive laws consider a negligible resistance ('buckling') of fibres under compression (Holzapfel et al, 2000; Federico and Herzog, 2008). Finite element (FE) models have been de- veloped for families of fibres oriented in different directions (Holzapfel et al, 2000; Holzapfel and Gasser, 2001). The constitutive approach based on angular integrals requires extensive calculations (Ateshian et al, 2009). Based on a hyperelastic free- energy function, these models have been improved using a generalized structure tensor that characterizes the statistical dispersed collagen fibre orientation (Gasser et al, 2006). Still, even if this approach requires a small number of calculations to obtain the strain and stresses fields, these models are limited (Cortes et al, 2010). Federico and Herzog (2008) proposed a constitutive model represented by an integral form of the elastic strain energy potential, which is performed on the unit sphere. Here, the integrals evaluation is performed using the spherical designs methods (Pierce et al, 2015). Recent progress in FE models has improved our understanding of depth-dependent properties of cartilage incorporating the fibre response using a reference direction and a fibre dispersion around this direction (Pierce et al, 2013; Wilson et al, 2007). Theoretical models to predict the colla- gen network architecture in various soft tissues assumes that collagen fibres align along the preferred fibre directions that are situated between the positive principal strain directions (Driessen et al, 2003; Wilson et al, 2006; Driessen et al, 2008). The collagen alignment has been shown to align with respect to loading in articu- Title Suppressed Due to Excessive Length 3 lar cartilage, predicting the development of the typical Benninghoff-type collagen fibre orientation (Wilson et al, 2004; Khoshgoftar et al, 2011). 1.1 Aim of the study In the present study, a novel continuum anisotropic hyperelastic formulation taking into account both fibres reorientation and fibres remodelling is proposed, aiming to investigate the role and evolution of the depth dependent collagen network in the articular cartilage. A structural tensor associated with the dispersion of the embedded collagen fibres is introduced. The arrangement of the fibres is not repre- sented by a numerical integration on the sphere surface (with a spherical harmonic distribution), but with a distribution of fibres simply defined around a reference direction and an ellipsoidal distribution assumed. The aim is to evaluate whether the new remodelling algorithm based on Wilson et al (2006), associated with this fibre-reinforced FE model is consistent with the orientation and distribution of the collagen fibres observed in native articular cartilage. The remodelling algorithm is therefore envisaged to be a valuable tool for the development of improved loading protocols for the tissue engineering of articular cartilage, and the understanding of structural adaptation of the collagen network during random loading conditions. 2 Continuum mechanical framework 2.1 Basic Kinematics Let Ω0 ⊂ R3 be the (fixed) reference configuration of a continuous body. The body undergoes a deformation χ, which transforms a typical reference material point X ∈ Ω0 into a spatial point x = χ (X) ∈ Ω in the deformed configuration. Let F(X) = ∂χ (X) /∂X be the deformation gradient and J (X) = det F(X) the local volumetric deformation ratio. Due to the material incompressibility of soft biological tissues (such as articular cartilage), the deformation gradient can be decomposed into a spherical (dilatational) and a unimodular (distortional) part as: (cid:16) (cid:17) ¯F F = J 1/3I (1) where I is the second order unity tensor and ¯F is associated to the part of the total deformation gradient that does not produce any change of volume. Soft tis- Fig. 1 Three-dimensional graphical representation of the ellipsoidal distribution of fibres given a reference direction ef,0. 4 S. Cortez et al. sues (such as cartilage) have a non-linear elastic mechanical behaviour. Generally, soft tissues like cartilage are modelled with pseudo-elastic or hyperelastic models. Adopting a hyperelastic formulation, the Lagrangian second Piola-Kirchhoff stress tensor (Π) can be derived from the strain energy density potential as Π = ∂W ∂E = 2 ∂W (C) ∂C (2) where C = FTF = J 2/3 ¯C and E = 1 tensor and the Green-Lagrange strain tensor, respectively. The scalar strain energy function (W ) depends on a set of invariants of C. This potential can also be decomposed into volumetric (Wv) and isochoric ( ¯W ) parts, 2 (C − I) are the right Cauchy-Green strain W(cid:0)J, ¯C(cid:1) = Wv (J) + ¯W(cid:0) ¯C, ...(cid:1) being the first term associated with the volumetric change and the second one with the isochoric deformation (Alves et al, 2010; Castro et al, 2014). In case of an inhomogeneous structure like articular cartilage, the mechanical response is driven by either the matrix or, mainly, by the collagen fibres. While the matrix determines the isotropic response, the collagen fibres determine the anisotropic response of the articular tissue. Therefore, the total elastic strain en- ergy can be decomposed into the elastic energy associated to the deformation of the matrix (i.e. the non-fibrous isotropic part) plus the elastic energy associated to the deformation of the collagen fibres (i.e. the fibrous anisotropic part), such that: ¯W(cid:0) ¯C, H(cid:1) = ¯Wiso In the present work, the isotropic material is described by the neo-Hookean (cid:0) ¯C, H(cid:1) (cid:0) ¯C(cid:1) + ¯Waniso (cid:0) ¯I1 − 3(cid:1) ¯Wiso( ¯C) = law, (3) (4) (5) (6) µ 2 (cid:104) and the anisotropic fibrous material by a new fibre-reinforced model proposed in this work, based on Gasser et al (2006) and Holzapfel and Ogden (2010), and on a new structure tensor H, ¯Waniso(C, H) = k1 2k2 exp(k2(cid:104)E1(cid:105)2) − 1 (cid:105) where µ is the shear modulus of the isotropic matrix of the articular cartilage, ¯I1 is the first invariant of C, H is the general structure tensor, and k1 and k2 are ma- terial parameters associated with fibres, which can be determined by experimental validation. E1 is a new pseudo-invariant defined as E1 = H : C − H : I (7) The new general structure tensor H plays a major role in the definition of this new pseudo-invariant E1. Indeed, the role of the structure tensor is to take into account any physically based distribution of the collagen fibres around a reference direction ef,0, as can be seen in the different layers of the articular cartilage. Therefore, the structure tensor H depends on b parameter, which defines Title Suppressed Due to Excessive Length 5 the dispersion of fibres around a random reference direction ef,0. The compact form for structure tensor can be given by: − H with b ∈ [−1, 0[, = 1 + β −bα(cid:17) (cid:16) I − bα(cid:16) H+ = (1 − bα) I + bα(cid:16) 1 + β 1 + β+(cid:17) −(cid:17) ef,0 ⊗ ef,0 ef,0 ⊗ ef,0 (8) (9) with b ∈ [0, 1]. The parameters α and β are material constants, varying between [−∞, +∞], aimed at increasing the flexibility of the model in order to allow a best fit of the nonlinear structure tensor stress response depending on parameter b to experimental data. In this new formulation, the range of b parameter is defined between −1 and 1. The following well-defined particular cases can be identified: a) lower limit b = −1 describes an isotropic distribution in the plane normal of the reference direction ef,0 (i.e., the superficial zone of the articular cartilage); b) the parameter b = 0 corresponds to the three-dimensional isotropic arrangement of the fibrous network (i.e., the middle layers of the articular cartilage); and c) the upper limit defines the alignment of all collagen fibres with the reference direction ef,0 (i.e., the deep zone). Fig. 1 shows a graphical representation of the fibres distribution with a particular reference direction ef,0. Table 1 shows the values of the distribution parameter b, a schematic representation of the orientation of fibres and the struc- ture tensor for each typical zone of the articular cartilage assuming the reference direction given by ef,0 ≡ z. A final remark concerning the usage of the Macaulay brackets (defined by the operator (cid:104)·(cid:105)) in Eq. 6. A common assumption is that fibres can only be loaded in tension, and suffer buckling in compression (Federico and Herzog, 2008). In order to take this into account, the elastic energy associated to the compression of the fibres must be neglected, and thus if E1 ≤ 0, the term inside the Macaulay brackets becomes zero and the energy function is reduced to the isotropic part. 2.2 Collagen remodelling The remodelling of the collagen network in the articular cartilage in general, and particularly in tissue engineered, occurs at two levels: on the one hand, the evo- lution of the collagen density, which determines the strength of the tissue; on the other hand, the reorientation of the collagen fibres network, which determines the anisotropy of the tissue. Together, density and orientation, as well as remodelling and reorientation, determine the mechanical response of the tissue to arbitrary mechanical loads. In what follows is presented the model proposed to describe both phenomena, i.e. the collagen fibres remodelling and reorientation. Knowing that collagen fibres growth and remodelling is due to the mechanical stimuli, and thus due to tissue deformation and strain fields, let us consider that e1 > e2 > e3, the principal vectors (eigenvectors), or principal strain directions, of the Green- Lagrange strain tensor E, and λ1 > λ2 > λ3 the principal values (eigenvalues) of the same strain tensor. The principal values can be either positive or negative, 6 S. Cortez et al. representing so a tensile or a compressive strain state, respectively, with respect to the correspondent principal strain vector (Driessen et al, 2003; Wilson et al, 2006; Khoshgoftar et al, 2011). Based on the hypothesis that collagen fibres align and growth along the positive principal strain directions, i.e. along the tensile mechanical stimulus, the fibre alignment is done through preferential direction ep (see Fig. 2) defined by: ep = e3 (10) when the principal value associated to a given principal vector is in compression and two principal values associated to the other two principal vectors are in tension (i.e. e1, e2 > 0 and e3 < 0) , or defined by ep = e1 (11) when two axes are in compression and an axis in tension (e1, e2 < 0 and e3 > 0). e1, e2 and e3 are the eigenvectors of the right-Cauchy tensor C, i.e., they are the principal directions of the strain. The collagen fibres tend to reorient toward the preferred fibre direction with a given angular velocity. Thus, the reorientation model defines the rate at which the reference axis ef,0 of the tissue rotates to the preferential direction ep: = καr = κ arccos(cid:107)ef,0 · ep(cid:107) dθ dt (12) where αr is the angle between the current reference fibre direction ef,0 (unde- formed configuration) and the preferred fibre direction ep determined at a given instant as a function of the local strain field (see Fig. 2). Fig. 2 Schematic representation of the preferred fibre direction ep situated in between the positive principal strain directions e1 and e2. Note that vector e3 is perpendicular to the plane. The fibre direction with respect to the undeformed configuration ef,0,old is rotated toward the preferred fibre direction ep over an angle dθ resulting in the new fibre direction ef,0. αr denotes the angle between ep and ef,0. To control the rate of reorientation, a positive constant κ was defined. The fibres are rotated around the following rotation axis, er = ef,0 ⊗ ep (cid:107)ef,0 ⊗ ep(cid:107) (13) However, future experimental results shall contribute to improve our understand- ing about fibre reorientation. Title Suppressed Due to Excessive Length 7 The collagen fibres will not only reorient but also redistribute around the refer- ence direction. The distribution of the collagen fibres around the reference direction is assumed to be described by parameter b and modelled by the structure tensor H. Adopting for parameter b a similar evolution law as adopted for the evolution of the reference direction, the following equation can be written as follow, = rb (bt − b0) db dt (14) where b0 is the current value of the distribution parameter b and bt is the target value of distribution of fibres, to be calculated from the strain field and strain loading history. The reorientation rate of distribution is defined by rb and it is associated with the regeneration process of the cartilage. To establish the natural phenomenological process of the tissue, we hypothesized that (i) in compression reaches the value of −1, i.e., the fibres tend to be oriented in the direction per- pendicular to the fibre reference direction, and (ii) in tension fibres tend to the positive value of +1, where the fibres are perfectly aligned with the fibre reference direction ef,0 (see Fig. 2). Table 1 Distribution parameter, schematic representation of the fibre orientation and struc- ture tensor for superficial, middle and deep zone of cartilage for a fibres reference direction of ef,0. Zone b Superficial -1.0 Middle 0.0 Deep 1.0 Fibre orientation Structure tensor in tension Structure tensor in compression 0 0 1 + β− bα  1 − bα  0 1 + β− bα 0 1 0 1   1 0 0 0 0   1 + β+bα 0 0 3 Finite element implementation The validation of the proposed formulation and FE implementation includes three main steps. Firstly, an analytical procedure to adjust the material and structure 8 S. Cortez et al. parameters was employed, determining a set of material parameters for each car- tilage zone. Using some literature studies (Ateshian et al, 1997; Elliott et al, 2002; Jurvelin et al, 2003; Pierce et al, 2013), the influence of these structure parameters were then investigated based on the nonlinear behaviour of the cartilage bipha- sic fibre-reinforced FE model. To explore the proposed remodelling algorithm, some numerical examples were implemented using the validated fibre-reinforced FE model, evaluating the performance of the collagen fibres reorientation that occurs in articular cartilage. At the end, a combination of all tissue layers in the same FE model was analysed. A total displacement corresponding to 10% of the specimen thickness was applied in the sample over a pseudo-time and different values of rb of 0.1, 0.2 and 0.3 (see Eq. 13) were evaluated. 3.1 Loading and boundary conditions Some studies and FE analysis in literature have been using a confined compression configuration to study the mechanical response of the collagen fibres in cartilage (Guo et al, 2015). However, this boundary value problem exhibits some main drawbacks in terms of the response of the fibres under tensile loads, i.e., the fibres strain in the transverse direction to the main loading axis (compressive loading) will be null due to the confinement. In the superficial zone, this issue is more evident and constraining. Thus, to demonstrate that the proposed formulation and FE model can reproduce the experimental observations, unconfined compression and uniaxial tension problems were considered. Both compression and tension Fig. 3 Cubic finite element mesh of a cartilage layer (1.0x1.0x1.0 mm3) symmetrically con- strained in x0y, x0z and y0z planes. , simulations were performed using a cubic FE mesh (1.0x1.0x1.0 mm3) discretized with 512 quadratic 27-node hexahedral finite elements (see Fig. 3). The FE model was symmetrically constrained in x0y, x0z and y0z planes and modelled on a home-developed open source FE solver. A fully-implicit Newton-Raphson iterative method to solve the nonlinear problem, and a biphasic formulation with total incompressibility of the solid matrix are used (Alves et al, 2010; Castro et al, 2014). An FE model of a cartilage cubic sample was built in unconfined compression for both superficial zone (with the fibres aligned parallel to articular surface) and middle zone (with the fibres randomly orientated). A displacement was applied to the upper surface of the sample up to an engineering strain of −0.50 and −0.80 for Title Suppressed Due to Excessive Length 9 the superficial zone and middle zone, respectively. As in the deep zone all fibres are in the same direction and perfectly aligned with the loading axis, the compressive response of the anisotropic part will be null due to the buckling assumption of the fibres under compression (Cortes et al, 2010). Thus, the results in compression of this zone were not considered for analysis. The uniaxial tension model was tested to validate the constitutive model with experimental curves for all cartilage zones (superficial, middle and deep). A displacement was applied to the upper surface of the sample up to an applied engineering strain of +0.15. In the superficial case, the loading axis is parallel to the plane in which all fibres are isotropically arranged, as represented in Fig. 4b, in opposition to the more biological orientation of the fibres with respect to the loading axis (see Fig. 4). In this tensile condition, fibres are stretched along the loading direction and compressed along the perpendicular directions. Here, a different fibre reference direction ef,0 was defined. Fig. 4 Orientation of the fibres in the superficial zone. Left: scheme of the fibres distribution in the plane perpendicular to the reference direction ef,0. The loading direction coincides with the reference direction. Right: scheme of the plane of fibre distribution aligned with the loading direction. Here, the reference direction of fibres was changed. 3.2 Material and structure data Analysed some studies from literature, the isotropic material properties of the ex- tracellular matrix were obtained. The shear modulus has been reported as constant for all zones in the articular cartilage. Despite the collagen content may influence the shear modulus (Pierce et al, 2013; Responte et al, 2007), in this study this collagen content was not considered in the definition of the material properties. The cartilage tension-compression nonlinear behaviour was analysed compar- ing the experimental data with the numerical results in order to find the structure parameters, i.e. parameters µ, k1, k2, alpha, beta+ and beta−. For the sake of simplicity, the same numerical model was used in compression and in tension to simulate the superficial and middle zones. However, for the deep zone, only the uniaxial tension model was achieved, given that fibres will not respond under com- pression when they are aligned with the loading direction. Unfortunately, there are no experimental data to compare the response of the deep zone of the articular cartilage. Here, the same material parameters used in the simulation of b = 0 were used for the simulation with b = 1, representing the deep zone when the fibres are aligned with the reference fibre direction. For all examples, ef,0 was defined 10 S. Cortez et al. to be the same as the loading direction, except for the uniaxial tension of the superficial zone, where it was assumed as being perpendicular to the loading di- rection (see Fig. 4). In this case, all fibres are dispersed in the perpendicular plane to the articular surface and the fibres are stretched in the loading direction. The fibre distribution parameter was defined as b = −1 for superficial zone (where the fibres are in the plane which is perpendicular to ef,0 and parallel to the articular surface), as b = 0 for the middle zone (with a randomly fibre distribution) and b = 1 for the deep zone (where all fibres are aligned with ef,0, i.e., perpendicular to the articular surface) (Wilson et al, 2006; Responte et al, 2007; Pierce et al, 2013). For each case, the structure parameters k1 and k2 were estimated. These Table 2 Reference material, structure and remodelling parameters of the FE model charac- terizing cartilage layers in different loading conditions (C - compression and T - tension). Zone Loading ef,0 b µ(MPa) k1(MPa) k2 α β+ β− bt κ Superficial Middle Deep C T C T T (0,0,1) -1.0 0.05 0.022 0.01 1.0 (1,0,0) -1.0 (0,0,1) 0.0 0.05 0.05 0.47 0.022 1.5 2.0 1.0 1.0 - - 1.0 (0,0,1) 0.0 0.05 0.53 0.1 1.0 1.0 (0,0,1) 1.0 0.05 0.135 0.01 1.0 1.0 0.1 -1.0 1.0 1.0 - - - - - 1.0 1.0 -1.0 1.0 1.0 1.0 rb 0.5 1.0 - 1.0 1.0 2.0 1.0 parameters are not well understood for cartilage (Pierce et al, 2013) but they may change with the depth of the tissue. The same is shown with α and β parameters. These new structure parameters introduced in this proposed model were defined as constant for all cases, except for compression in the superficial zone, where β was defined as 0.1 to adjust better the tension-compression curve to experimental data. A convergence problem in the numerical example led to use a different value of β. To better understand the influence of α and β structure parameters, some representative numerical examples were performed (see Sect. 4.1). Table 2 gives the set of material and structure parameters used in the proposed fibre-reinforced FE model to estimate the nonlinear response of the three cartilage layers (superficial, middle and deep zones) under compression (C) and tension (T) conditions. 3.3 Remodelling analysis The remodelling algorithm was integrated in the fibre-reinforced FE model. Some numerical examples using the proposed approach were performed for all zones of cartilage (superficial, middle and deep). To illustrate the major features of the fibre remodelling model, the cubic sample was loaded axially (in tension and com- pression) and the axial true stress was analysed. For each zone, the material and structure values were assumed the same as in the previous examples (see Table 2). Title Suppressed Due to Excessive Length 11 Initially, the remodelling parameter rb was arbitrary set to 1.0. In parallel with simulations, its influence in the fibre distribution parameter was analysed. As the value of the constant rate κ showed to be not critical for the evolution of the remodelling process and as there is no experimental data to validate this param- eter, it was set 1.0 for all simulations. The remodelling process was considered with a dimensionless time scale. The load was applied instantaneously, thereafter it is held constant and the remodelling process starts. The evolution of the fibre orientation and stress fields are analysed during the remodelling process, as shown in Fig. 14-16. The remodelling parameters used in all simulations are displayed in Table 2. 4 Results and Discussion 4.1 Tension-compression nonlinear behaviour After an extensive exploration for data which might validate our FE model, some literature data (Ateshian et al, 1997; Elliott et al, 2002; Jurvelin et al, 2003; Pierce et al, 2013) was selected. A uniaxial unconfined compression test was carried out to demonstrate the validation of the constitutive model for the superficial zone (with the fibres aligned in the loading direction). Fig. 5 shows a comparison between the computed axial Cauchy stress component obtained by the FE model and com- pared with the experimental results available (Ateshian et al, 1997). Simulation coincides well within one standard deviation of the stresses determined experimen- tally, reproducing the mechanical behaviour of the superficial zone (distribution defined as b = −1). When the cuboic FE model is loaded, the resulting axial stress increases (in absolute value) nonlinearly with the stretch. The FE model validation required a little effort. Although, a difference between a confined and unconfined compression was not considered in this study and it is believed that probably this different is associated with the hard validation. To achieve a better adjustment to the experimental results, and as the nonlinearity of the stress-strain curves can be controlled with the structural parameters, a different value of β parameter (0.1 instead of 1.0) was defined in this simulation. The analysis of β parameter is presented below. Fig. 5 Unconfined compression stress of the FE simulation for the superficial zone and com- parison with the experimental data from Ateshian et al (1997). 12 S. Cortez et al. An FE model of the middle cartilage zone with a random fibre orientation was study using the experimental data from Jurvelin et al (2003). Fig. 6 shows the computed axial true stress versus engineering strain for both simulation model and experimental data. The proposed fibre-reinforced FE model fits the nonlinear behaviour (in agreement with standard deviations from exper- imental results) and it is able to reproduce the experimental data in unconfined compression conditions. Fig. 6 Unconfined compression stress of the FE simulation for the middle zone and comparison with the experimental data from Jurvelin et al (2003). The uniaxial tension simulations of the superficial and middle zones are pre- sented in Fig. 7 and Fig. 8, respectively. In the first simulation, all fibres are defined in the plane parallel (perpendicular to the fibre reference direction) to the loading direction. The results exactly fit the experiments (Elliott et al, 2002) for tensile testing of this zone very accurately. Similarly, the tensile results for the middle zone overlap the experimental data from Elliott et al (2002). Fig. 7 Uniaxial tension stress of the FE simulation for superficial zone of cartilage and com- parison with the experimental data from Elliott et al (2002). To demonstrate that the proposed fibre reinforced FE model can be used to simulate the cartilage deep zone, a representative unconfined compression test (using the experimental data from the middle zone (Elliott et al, 2002) was per- formed. Fig. 9 shows the results of this representative simulation. Although the experimental data in this case are not available, the model proved to be suitable Title Suppressed Due to Excessive Length 13 Fig. 8 Uniaxial tension stress of the FE simulation for middle zone of cartilage and comparison with the experimental data from (Elliott et al, 2002). in cases where the fibres are perfectly aligned (b = 1). For feasible experimental data of the deep zone, the model parameters should be adjusted. Fig. 9 Uniaxial tension stress of the FE simulation for deep zone of cartilage and validation with data of the middle zone stress-strain curve from Elliott et al (2002). In this case, the parameters should be adjusted. Fig. 10 The influence of structure parameter on stress response for different fibres distribu- tions under compression and tension simulations. 14 S. Cortez et al. The role of structure parameters α and β The influence of α and β parameters in the tension-compression response was also investigated. These parameters are used to complement the structure tensor increasing the flexibility of the model. A representative numerical example was performed and the relationship between the fibre distribution (b) and the variation of the stress with the β parameter is plotted in Fig. 10. In all cases, the parameter α was kept constant as 1.0. For β (+ and −) defined as zero, the stress increases in tension and it is constant in compression for negative values of b (superficial zone) until reach the totally isotropic region (middle zone with b = 0). Consequently, for a positive value of b the behaviour is opposite, the axial stress is constant in tension and increases in compression (from the isotropic distribution to the distribution of perfectly fibres aligned). When β (+ and −) is considered in the structure tensor for values different from zero, stress increases in tension from the superficial zone (where b is more negative) to the deep zone (b = −1) (see Fig. 11). The opposite occurs when the model is under compression. If the β parameter increases (see Fig. 12), the maximum value of the stress will increase in compression and also in tension. Fig. 11 The influence of structure parameter β = 1 on stress response for different fibres distributions under compression and tension simulations. Fig. 12 The influence of structure parameter β = 2 on stress response for different fibres distributions under compression and tension simulations. Another parameter which characterizes the structure tensor is the parameter α, which allows the adjustment of the stress response curvature as function of the Title Suppressed Due to Excessive Length 15 fibre alignment. As shown in Fig. 13, the change of this parameter gives a different result in the stress response. Fig. 13 The influence of structure parameter α on the stress response for different fibres distributions under compression and tension simulations. 4.2 Remodelling process To evaluate the new remodelling approach, the evolution of the fibre distribu- tion parameter and the fibre reorientation under different loading conditions was investigated. Fig. 14−17 compares the evolution of b in unconfined compression versus pseudo-time stepping, with a different value for rb parameter. Both simu- lations show (Fig. 14) that the distribution of fibres tends to negative values, as it should, since that all fibres, initially aligned (b0 = 1) with the reference direction ef,0 = (1, 0, 0), tend to be distributed isotopically in the plane perpendicular to the loading direction, and the reference fibre vector becomes perpendicular to this reference direction. However, an increase of rb from 0.5 to 1.0 leads to the b param- eter reaches faster the expected value bt = −1. It should be noted that this case does not correspond to any point of the cartilage and it is only referred to a better understand the presented model. Fig. 15 shows the remodelling of b for a reference Fig. 14 Remodelling of fibre distribution versus time in compression with a reference fibre direction ef,0 = (1, 0, 0) and b0 = 1 (representing an unknown zone). The dotted line with a rb = 0.5 and the continuous line with rb = 1.0. 16 S. Cortez et al. fibre direction of ef,0 = (0, 0, 1), when fibres are organized following the plane perpendicular to this direction (b0 = −1). Thus, it represents the superficial zone of the cartilage. When this zone is compressed, the distribution of fibres (bt = −1) tends to be constant along the simulation. There is no remodelling process of fibres because in compression they will expand in the same plane and no change in the directions is observed. The same occurs when the distribution of fibres is aligned with the reference direction. The distribution was kept as constant (b = 1) during the simulation (see Fig. 15). This result can show the natural phenomenon that occurs with the fibres in the deep zone. When the tissue is loaded in tension, all fi- bres aligned with the reference direction tend to stretch (along the axial direction) and keeping the same direction while they are in tension. In compression, fibres in this layer of the tissue do not respond, resulting in a phenomenon called 'buckling' of fibres (Federico and Herzog, 2008; Cortes et al, 2010). Thereby, only the tension simulation was presented in this study. To analyse the evolution of the isotropic Fig. 15 Remodelling of fibre distribution versus time in compression with a reference fibre direction ef,0 = (0, 0, 1) and b0 = −1 (representing the superficial zone).Remodelling of fibre distribution versus time in tension with a reference fibre direction ef,0 = (0, 0, 1) and b0 = 1 (representing the deep zone). distribution, which is associated with the middle zone (b0 = 0) of the tissue, one model in tension and two models in unconfined compression were investigated. The fibre reference direction was defined by ef,0 = (0, 0, 1) in both simulations. The results of these three simulations are presented in Fig. 16. In tension, b0 is Fig. 16 Remodelling of fibre distribution versus time in compression and tension with a reference fibre direction ef,0 = (0, 0, 1) and b0 = 0 (representing the middle zone). Title Suppressed Due to Excessive Length 17 initialized as zero and tends to positive values (where the limit is bt = 1). This reorientation of fibres indicates that, in tension, all randomly oriented fibres tend to realignment with the loading direction, which is the same of the reference fibre direction. In compression, the distribution of fibres varied between b0 = 0 and bt = −1. Initially, fibres show an isotropic distribution (middle zone) and then, they tend to be perpendicular to the reference direction ef,0 = (0, 0, 1) but with an isotropic distribution in the plane (bt = −1). As shown in a previous numerical example (Fig. 14), when the rb increases the negative limit (bt = −1) is reached faster (rb = 2). These results highlighted the assumptions initially made in this work (see Sect. 2.1). In compression the distribution of fibres parameter tends to be −1, characterizing a plane isotropic distribution and in tension the distribu- tion parameter tends to be 1, describing the perfect alignment of fibres with the reference direction (Driessen et al, 2003; Wilson et al, 2006). FE model with three layers To explore the influence of the distribution of fibres parameter (b) under deformation with a depth-dependent manner, a combination of three layers in the same FE model was performed. The FE mesh used previ- ously was divided into three different materials with 20%, 50% and 30% of the total thickness for superficial, middle and deep zone, respectively. The fibre distribution for each zone was defined as b = −1.0 (superficial), b = 0.0 (middle) and b = 1.0 (deep) with a fibre reference direction ef,0 = (0, 0, 1). However, these b values will change as function of tissue deformation history. The FE model was compressed by 10% of the sample thickness over a pseudo-time. The axial displacement was applied perpendicular to the top of superficial zone. The results of these simula- tions are presented in Fig.17 with different values for the remodelling parameter a)rb = 0.1 b)rb = 0.2 and c)rb = 0.3. For the same time step, the remodelling pa- Fig. 17 Compression simulation of a three-layer FE model with different fibres distribution and three different remodelling parameters a)rb = 0.1 b)rb = 0.2 and c)rb = 0.3. rameter (rb) influences the distribution of fibres (b). With deformation, fibres run to the direction which can resist more. As previously discussed, when the sample is compressed, fibres reorient to have a distribution of b = −1.0. This phenomenon is evident in the middle and deep zones, while in the superficial zone it keeps constant along the time. The remodelling parameter controls the rate of change of the distribution of fibres. When the rb value increases, the remodelling process is faster to the new direction, determined from the deformation history. It is shown Fig.17 where the distribution in the deep zone is b ≈ 0.8 for rb = 0.1 (Fig.17a), b ≈ 0.6 for rb = 0.2 (Fig.17b) and b ≈ 0.3 for rb = 0.3 (Fig.17c). 18 5 Conclusion S. Cortez et al. In the present study, a novel continuum anisotropic hyperelastic remodelling ap- proach to investigate the effects of the cartilage collagen network in a depth- dependent manner was proposed. A new structure tensor, in order to describe the mechanical response of the embedded anisotropic collagen fibres, was intro- duced and different material and structure parameters were included to increase the flexibility of the new proposed model. The preliminary numerical results al- low to demonstrate the ability of the new formulation to reproduce the nonlinear tension-compression response of cartilage in two main loading conditions, and cross validated with the experimental data available from the literature. In the super- ficial zone, the validation of these parameters was more complex either by the absence of experimental data or by the confined configuration used in the current experiments. In case of the shear modulus of the isotropic matrix, the value was assumed as equal in all regions being low when compared with literature. The proposed FE model showed good results and low numerical consumption when compared with other approaches in literature. This formulation is relatively simple when compared with the numerical integration on the unit sphere (Federico and Herzog, 2008) where the spherical designs methods are used. In the present model the dispersion of fibres was considered in the hyperelastic formulation through a structure tensor, which gives a different mechanical response when compared with models where the anisotropic strain depends only by the local reference direction of the collagen fibres (Holzapfel and Ogden, 2010; Pierce et al, 2013, 2015). More- over, the buckling phenomenon of the fibres was also taken into account. In the remodelling cases, the evolution of the collagen fibres orientation for each zone was analysed. Again, due to the lack on experimental data, the study of re- generation fibre parameter was not totally achieved. Thus, some remodelling pa- rameters were fixed in all simulations. Although this work has focused on changes in fibres orientation, the effects of fibre remodelling in the mechanical properties of the tissue should be further investigated. Diverse fibre distributions could be explored using this model and more real-like boundary conditions used instead of the layer-based sample. This new constitutive law for modelling the anisotropy, reorientation and remod- elling behaviour of the articular cartilage can contribute to the development of the future studies in the articular cartilage. In addition to cartilage tissue, the present model has the potential to study other fibrous soft tissues, contributing to a better knowledge of the structural adaptation in the collagen network associated with arbitrary mechanical stimuli. In summary, this study aims to improve nu- merical models that have proven to be essential tools to support the experimental protocols in the cartilage tissue engineering. 6 Conflict of interest The authors declare that they have no conflict of interest. Acknowledgements The first author is grateful to FCT -- Funda¸cao para a Ciencia e a Tecnologia (Portugal) for the PhD grant (SFRH/BD/87933/2012). This work is funded by FEDER through COMPETE - Programa Operacional de Fatores de Competitividade (COMPETE 2020- Title Suppressed Due to Excessive Length 19 PTDC/EMS-TEC/3263/2014) and by national funds through FCT, under the strategic project PEst-C/EME/UI0481/2013 and also in the scope of the following project: FCOMP-01-0124- FEDER-015191. References Alves J, Yamamura N, Oda T, Teodosiu C, Middleton J, Evans S, Holt C, Jacobs C, Atienza C, Walker B (2010) Numerical simulation of musculo-skeletal systems by v-biomech. Proceedings of 9th International Symposium CMBBE2010 Ateshian G, Warden W, Kim J, Grelsamer R, Mow V (1997) Finite deformation biphasic material properties of bovine articular cartilage from confined compres- sion experiments. Journal of biomechanics 30(11):1157 -- 1164 Ateshian GA, Rajan V, Chahine NO, Canal CE, Hung CT (2009) Modeling the matrix of articular cartilage using a continuous fiber angular distribu- tion predicts many observed phenomena. Journal of biomechanical engineering 131(6):061,003 Bandeiras C, Completo A, Ramos A (2015) Influence of the scaffold geometry on the spatial and temporal evolution of the mechanical properties of tissue- engineered cartilage: insights from a mathematical model. Biomechanics and modeling in mechanobiology pp 1 -- 14 Castro A, Wilson W, Huyghe J, Ito K, Alves J (2014) Intervertebral disc creep behavior assessment through an open source finite element solver. Journal of biomechanics 47(1):297 -- 301 Chung C, Ho SY (2010) Analysis of collagen and glucose modulated cell growth within tissue engineered scaffolds. Annals of biomedical engineering 38(4):1655 -- 1663 Cortes DH, Lake SP, Kadlowec JA, Soslowsky LJ, Elliott DM (2010) Charac- terizing the mechanical contribution of fiber angular distribution in connective tissue: comparison of two modeling approaches. Biomechanics and modeling in mechanobiology 9(5):651 -- 658 Driessen N, Peters G, Huyghe J, Bouten C, Baaijens F (2003) Remodelling of continuously distributed collagen fibres in soft connective tissues. Journal of biomechanics 36(8):1151 -- 1158 Driessen NJ, Cox MA, Bouten CV, Baaijens FP (2008) Remodelling of the angular collagen fiber distribution in cardiovascular tissues. Biomechanics and modeling in mechanobiology 7(2):93 -- 103 Elliott DM, Narmoneva DA, Setton LA (2002) Direct measurement of the poisson ratio of human patella cartilage in tension. Journal of biomechanical engineering 124(2):223 -- 228 Federico S, Gasser TC (2010) Nonlinear elasticity of biological tissues with statis- tical fibre orientation. Journal of the Royal Society Interface 7(47):955 -- 966 Federico S, Herzog W (2008) Towards an analytical model of soft biological tissues. Journal of biomechanics 41(16):3309 -- 3313 Gasser TC, Ogden RW, Holzapfel GA (2006) Hyperelastic modelling of arterial layers with distributed collagen fibre orientations. Journal of the royal society interface 3(6):15 -- 35 20 S. Cortez et al. Guo H, Maher SA, Torzilli PA (2015) A biphasic finite element study on the role of the articular cartilage superficial zone in confined compression. Journal of biomechanics 48(1):166 -- 170 Holzapfel GA, Gasser TC (2001) A viscoelastic model for fiber-reinforced compos- ites at finite strains: Continuum basis, computational aspects and applications. Computer methods in applied mechanics and engineering 190(34):4379 -- 4403 Holzapfel GA, Ogden RW (2010) Constitutive modelling of arteries. Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences 466(2118):1551 -- 1597 Holzapfel GA, Gasser TC, Ogden RW (2000) A new constitutive framework for arterial wall mechanics and a comparative study of material models. Journal of elasticity and the physical science of solids 61(1-3):1 -- 48 Hossain MS, Bergstrom D, Chen X (2014) Prediction of cell growth rate over scaffold strands inside a perfusion bioreactor. Biomechanics and modeling in mechanobiology 14(2):333 -- 344 Jurvelin J, Buschmann M, Hunziker E (2003) Mechanical anisotropy of the human knee articular cartilage in compression. Proceedings of the Institution of Me- chanical Engineers, Part H: Journal of Engineering in Medicine 217(3):215 -- 219 Khoshgoftar M, van Donkelaar CC, Ito K (2011) Mechanical stimulation to stim- ulate formation of a physiological collagen architecture in tissue-engineered car- tilage: a numerical study. Computer methods in biomechanics and biomedical engineering 14(02):135 -- 144 Khoshgoftar M, Wilson W, Ito K, van Donkelaar CC (2013) The effect of tissue- engineered cartilage biomechanical and biochemical properties on its post- implantation mechanical behavior. Biomechanics and modeling in mechanobiol- ogy 12(1):43 -- 54 Nava MM, Raimondi MT, Pietrabissa R (2013) A multiphysics 3d model of tissue growth under interstitial perfusion in a tissue-engineering bioreactor. Biome- chanics and modeling in mechanobiology 12(6):1169 -- 1179 Pearle AD, Warren RF, Rodeo SA (2005) Basic science of articular cartilage and osteoarthritis. Clinics in sports medicine 24(1):1 -- 12 Pierce DM, Ricken T, Holzapfel GA (2013) A hyperelastic biphasic fibre-reinforced model of articular cartilage considering distributed collagen fibre orientations: continuum basis, computational aspects and applications. Computer methods in biomechanics and biomedical engineering 16(12):1344 -- 1361 Pierce DM, Unterberger MJ, Trobin W, Ricken T, Holzapfel GA (2015) A mi- crostructurally based continuum model of cartilage viscoelasticity and perme- ability incorporating measured statistical fiber orientations. Biomechanics and modeling in mechanobiology pp 1 -- 16 Responte DJ, Natoli RM, Athanasiou KA (2007) Collagens of articular cartilage: structure, function, and importance in tissue engineering. Critical Reviews in Biomedical Engineering 35(5) Wilson W, Van Donkelaar C, Van Rietbergen B, Ito K, Huiskes R (2004) Stresses in the local collagen network of articular cartilage: a poroviscoelastic fibril- reinforced finite element study. Journal of biomechanics 37(3):357 -- 366 Wilson W, Driessen N, Van Donkelaar C, Ito K (2006) Prediction of collagen ori- entation in articular cartilage by a collagen remodeling algorithm. Osteoarthritis and Cartilage 14(11):1196 -- 1202 Title Suppressed Due to Excessive Length 21 Wilson W, Huyghe J, Van Donkelaar C (2007) Depth-dependent compressive equi- librium properties of articular cartilage explained by its composition. Biome- chanics and modeling in mechanobiology 6(1-2):43 -- 53
1606.01195
2
1606
2017-03-02T16:54:45
A Flexible Phase Retrieval Framework for Flux-limited Coherent X-Ray Imaging
[ "physics.bio-ph" ]
Coherent X-ray diffraction imaging (CXDI) experiments are intrinsically limited by shot noise, a lack of prior knowledge about the sample's support, and missing measurements due to the experimental geometry. We propose a flexible, iterative phase retrieval framework that allows for accurate modeling of Gaussian or Poissonian noise statistics, modified support updates, regularization of reconstructed signals, and handling of missing data in the observations. The proposed method is efficiently solved using alternating direction method of multipliers (ADMM) and is demonstrated to consistently outperform state-of-the-art algorithms for low-photon phase retrieval from CXDI experiments, both for simulated diffraction patterns and for experimental measurements.
physics.bio-ph
physics
A Flexible Phase Retrieval Framework for Flux-limited Coherent X-Ray Imaging Department of Electrical Engineering, Stanford University, Stanford, California 94305, USA Liang Shi and Gordon Wetzstein SLAC National Accelerator Laboratory, Menlo Park, California 94025, USA∗ (Dated: March 3, 2017) Thomas J. Lane* Coherent X-ray diffraction imaging (CXDI) experiments are intrinsically limited by shot noise, a lack of knowledge about the sample's support, and missing measurements due to the experimen- tal geometry. We propose a flexible, iterative phase retrieval framework that allows for accurate modeling of Gaussian or Poissonian noise statistics, modified support updates, regularization of reconstructed signals, and handling of missing data in the observations. The proposed method is efficiently solved using alternating direction method of multipliers (ADMM) and is demonstrated to consistently outperform state-of-the-art algorithms for low-photon phase retrieval from CXDI experiments, both for simulated diffraction patterns and for experimental measurements. Propelled by the development of ultra-bright 4th generation light sources like the linac coherent light source coherent X-ray diffraction imag- ing (CXDI) has the promise of revealing the atomic or near-atomic structure of aperiodic structures like single, non-crystalline proteins [1–4]. (LCLS), In a common implementation of such experiments, identically-structured particles are exposed to coherent X-ray pulses at unknown orientations and a detector records the intensity of forward-scattered diffraction pat- terns. Real-space images can then be retrieved by means of iterative phase retrieval algorithms [5–9], for exam- ple the hybrid input-output (HIO) and relaxed averaged alternating reflections (RAAR). To employ any of these algorithms, it is necessary to "oversample" the diffraction patterns by a factor of at least 2x (in the Shannon sense) [10]. Due to this oversam- pling, reconstructed molecules only occupy part of the image, known as the support. Unfortunately, the specific support of any sample is usually unknown and has to be estimated. For example, the Shrinkwrap algorithm [2], one of the most common approaches, thresholds the real- space image at each iteration to provide a continuously updated support estimate. An accurate support is cru- cial for reliable reconstructions. A second unavoidable challenge in phase retrieval is Poisson-distributed shot noise due to the fact that source brightness or radiation damage intrinsically limits the number of photons a given sample can diffract [11]. Re- construction resolution is determined by the total num- ber of photons diffracted [2]. Most iterative phase re- trieval algorithms implicitly employ a Gaussian noise model (either as additive readout noise or as a model of the diffraction intensity), which is not optimal for low- light imaging. This model mismatch is small when many photons are captured by each camera pixel (≫ 10), but it becomes significant when photons are scarce. Modeling Poisson noise directly with maximum-likelihood methods has demonstrated reconstruction improvement in pty- chography [12] and aberration estimation in incoherent imaging [13], suggesting it may also offer advantages in CXDI reconstructions. Finally, in most CXDI experiments, the direct beam necessitates a beam dump or stop to prevent damage to experimental equipment, resulting in a missing re- gion in the center of the detector that corresponds to low frequency information. This missing information can severely limit the quality of a reconstruction or make it impossible [3, 14–16]. Since missing frequencies are un- constrained by available experimental data, they often remain near the random initial values set by the phase retrieval method employed, resulting in significant ar- tifacts in the reconstructed image. A common empiri- cal solution is to fix the missing intensities to physically motivated values (see e.g. [3]). Recent studies [17, 18], however, demonstrated that prior information about the real space image, such as total variation regularization (TV) [19], could sufficiently constrain missing intensi- ties, resulting in reproducible, high-quality reconstruc- tions. However, the regularization was carried out in an ad hoc manner separate from the phase retrieval algo- rithm employed (HIO) and therefore was not easily in- terfaced with other methods. In this letter, we propose a framework for combining different prior information in an efficient and flexible way, leveraging the alternating direction method of multipli- ers (ADMM) [20]. Similar approaches have been shown successful in phase retrieval for sparse signals [21, 22]. Within this framework, we implement a Poisson error model, TV regularization, support update model, and handle missing measurements in a unified manner. Our method is compatible with classic update rules like HIO and RAAR. Here, we focus on the most common case of real and positive images, in both real and diffraction space; extensions to e.g. complex images is straightfor- ward. To be precise, we introduce a notation for CXDI mea- surements and quickly review these classic phase retrieval algorithms before describing our method mathematically. Let x be a finite scalar field representing the electron den- sity of an object, generally in 3 dimensions. In this paper, we limit our interest to 2D manifolds within a 3D object and 2D objects. Extension to 3D objects is straight for- ward. We employ a vectorized representation x ∈ RN , where xi denotes a voxel value at index i, and i runs over all voxels in all dimensions. To deal with missing data in the diffraction image (due to e.g. the beam stop), let w ∈ RN be a binary vector with value 0 for pixels in missing regions and 1 otherwise. In CXDI, the camera measures the diffraction image (a) (b) b = F x2, where b ∈ RN is the measured Fourier intensity and F is the 3D discrete Fourier transform operator. Let Sk be the support, that is xi = 0 if i /∈ Sk, at iteration k. Define projection operator Pm, Pm(x) = F −1(v), vi =   which enforces the amplitude of the diffraction image from the estimated object to equalize the experimental measurement. The Fourier components in missing re- gions remain unchanged [23]. Wi = 1, F x 6= 0 otherwise pbi (F x)i (F x)i (F x)i The three classic algorithms we consider consist of it- erative updates, where in-support pixels are updated ac- cording to xk+1 i = (Pmxk)i, i ∈ Sk, and out-of-support pixels (i /∈ Sk) are updated by one of (where the algorithm is indicated) xk+1 i =   Error Reduction 0 xk i − β(Pmxk)i HIO i + (1 − 2β)(Pmxk)i RAAR βxk with β a feedback parameter (a typical choice of β is around 0.9 [8]). The update rule for pixels inside the support can be interpreted as a gradient descent update 2, followed by projection onto the feasible set in real space, i.e. support Sk [9]. which minimizes εm = kdiag(w)(F x−√b)k2 The model above assumes the diffracted wavefront can be measured perfectly. A real experiment is better mod- eled by a Poisson process that takes into account photon shot statistics b ∼ P(cid:0)F x2(cid:1) . specifically, the probability of observing bi photons at pixel i is p (bix) =   (cid:0)F x2(cid:1)bi i e−(F x2)i bi! const , Wi = 1 , Wi = 0 42.6dB 19.7dB 2 Ground Truth ADMM Result 31.5dB ADMM Support 23.8dB Hawk Result Hawk Support 23.8dB Ground Truth -5 ADMM Result HIO+ER Result Hawk Result B d n i E S M R e c a p S ) l a e R ( e g a m I -10 -15 -20 -25 -30 -35 0 HIO+ER ADMM(HIO+ER) Hawk+ER (c) 200 400 600 800 1000 1200 1400 1600 1800 2000 Iteration FIG. 1. (a) Reconstructed "Cameraman" and associated sup- ports from 2.5x oversampled noise-free simulations, for both our method (ADMM) and the Hawk package using σinit = 3, σend = 1.5. PSNRs labeled on top-left corner of each image. (b) Reconstructed "Lena" from 2x oversampled simulations with shot-noise (average 150 photons/px), for our method (ADMM), HIO and ER implemented by the authors, and the Hawk package using σinit = 1, σend = 0.5. PSNRs labeled on top-left corner of each image. (c) Real-space root-mean- square error (RMSE) of "Lena" between the ground truth and reconstruction over 2000 iterations. Let 1 be a column vector with every element 1, the log- likelihood of the joint probability can then be expressed in following compact form LW (x) = (cid:0)W logF x2(cid:1)T b−(cid:0)WF x2(cid:1)T where W = diag(w), with gradient N Xi=1 1− Wilog (bi!) ∇LW (x) = 2F −1(cid:18)F x−W(cid:16)diag(F x2)−1diag(F x)b(cid:17)(cid:19) allowing the log-likelihood to be efficiently maximized by gradient ascent. We add two pieces of prior knowledge to this model of the measurement: that we expect the real space im- age to be smooth and positive. This is not an exhaus- tive list of prior information that could be employed, but will be used to demonstrate the effectiveness of our approach. We add prior terms directly to the ob- jective function, allowing for continuous enforcement of this information. No intensity constraints are applied to missing-measurement regions, though these regions are implicitly constrained by the real-space priors. We can then write an objective function for the phase retrieval problem with these priors, minimize x − LW (x) + λkDxk2,1 + IR+ (x) , (1) where D = [Dx; Dy] ∈ R2N ×N is the discrete gradi- ent operator consisting the horizontal and vertical par- tial derivative operators(vertically stacked), kDxk2,1 = Pg∈{x,y} kDgxk2 is the ℓ1/ℓ2 norm of Dx, λ is the weight for the isotropic TV term and IR+ (x) is the indicator function enforcing positivity (IR+ = 0 if xi ∈ R+ and ∞ otherwise). During iteration, IR+ is not evaluated in the objective, only in the gradient. This problem can be efficiently solved by the alternat- ing direction of multipler methods (ADMM) [20]. Using ADMM, Eq. (1) is reformulated as minimize x subject to f (x) −LW {z } (cid:20) D I (cid:21) {z }K g1(z1) + λkz1k2,1 {z } z2 (cid:21) x −(cid:20) z1 {z }z + IR+(z2) {z } g2(z2) = 0, (2) (3) where z1 ∈ R2N and z2 ∈ RN are slack variables. ADMM splits the objective into a weighted sum of three inde- pendent functions f (x), g1(z1) and g2(z2) that are only linked through the stated constraints. Following the gen- eral ADMM strategy, we write an augmented Lagrangian of Eq. (2) Lρ1,ρ2 (x, z, y) =f (x) + g1 (z1) + g2 (z2) + yT (Kx − z) 2 , (4) ρ1 2 kDx − z1k2 ρ2 2 kx − z2k2 2 + + where ρ1 and ρ2 set the penalty associated with ADMM constraints violation. Under the scaled form of the aug- mented Lagrangian, a single ADMM iteration consists of a sequential updates: f (x) + x x ← proxquad,ρ1,ρ2 (v1, v2) = arg min ρ1 2 kDx − v1k2 v1 = z1 − u1, v2 = z2 − u2 z1 ← proxk·k1,ρ1 (v) = arg min ρ2 2 kx − v2k2 2 , 2 + z1 g1 (z1) + ρ1 2 kv − z1k2 2 , v = Dx + u1 z2 ← proxI,ρ2 (v) = arg min z2 g2 (z2) + ρ2 2 kv − z2k2 2 , v = x + u2 ← u + Kx − z, (5) u2 (cid:21) (cid:20) u1 {z }u where the scaled dual variable u = (1/ρ)y is used to simplify the notation of Eq. (4). (cid:18) (cid:20)(cid:22)(cid:15)(cid:18)(cid:69)(cid:35) (cid:19)(cid:20)(cid:15)(cid:18)(cid:69)(cid:35) (cid:19)(cid:19)(cid:15)(cid:17)(cid:69)(cid:35) (cid:18)(cid:25)(cid:15)(cid:26)(cid:69)(cid:35) 3 (cid:17) (cid:34)(cid:37)(cid:46)(cid:46) (cid:41)(cid:66)(cid:88)(cid:76) (cid:34)(cid:37)(cid:46)(cid:46) (cid:41)(cid:66)(cid:88)(cid:76) Ground Truth 12.5 photons/px 1.25 photons/px FIG. 2. Reconstructed caffeine molecule from 2x oversam- pled simulations with shot-noise at average [12.5, 1.25] pho- tons/px, for our method (ADMM) and the Hawk package us- ing σinit = 1, σend = 0.5. PSNRs labeled on top-right corner of each image. The x-update is a quadratic program which can be iteratively minimized by gradient-based meth- ods. We used the Hessian-free Newton method pro- vided by the minFunc package (http://www.cs.ubc.ca/ schmidtm/Software/minFunc.html). Upon completion, out-of-support pixels can be updated using rules speci- fied by HIO or RAAR. After t ADMM iterations, we update the object sup- port using a modified Shrinkwrap algorithm. Let G(kr) be a normalized Gaussian blurring kernel with standard deviation kr, and η be a thresholding parameter. Stan- dard Shrinkwrap updates the support by convolving the realspace image with a Gaussian and thresholding, pre- cisely 1. xg = x ∗ G(kr) 2. S = {i (xg)i ≥ η max(xg)}. Inspired by recent study which showed that modest over- estimation of the extent of the support results in signifi- cantly less error than underestimating it [16], we added a third step that fills in any non-support regions completely enclosed by the support. Specifically, we employed a mor- phological hole-filling on the binary support image [24]. We also experimented with including the entire convex hull of the Shrinkwrap support, with good but inferior results. For the z1 and z2 update steps, closed-form solutions can be obtained using proximate operators for the ℓ1/ℓ2- norm and indicator function, as discussed in [25]. The ADMM framework allows flexible expression of models that can be customized easily and solved effi- ciently. For example, switching to a Gaussian error model only requires modification on the cost function of the x- update and its gradient. Because the slack variable (z) update is separated from the unknown variable (x) up- date, changing or imposing new priors is convenient. To assess the proposed algorithm, we applied it to both simulated diffraction patterns and experimental measurement imaged at LCLS. The results are com- pared with the ones produced by the state-of-the-art phase retrieval toolbox Hawk [26], which implements the standard shrinkwrap algorithm, HIO and RAAR with a Gaussian noise model (but not a Poisson noise model). We began with test images (Figures 1 and 2). Ini- tial supports were obtained by thresholding the inverse Fourier transform of intensity measurement (autocorre- lation) at 4% of the maximum intensity. The initial and minimal standard deviation of Shrinkwrap Gaus- sian blur kernel(σinit, σmin) were chosen based on trails and the initial value was reduced by 1% every 20 it- erations down to the minimum. The parameters η = [5%, 10%, 15%, 20%] and λ = [0.1, 0.01, 0.005, 0.001] were tested and the resulting highest peak signal-to-noise ratio (PSNR) reconstruction was chosen. The slack vari- ables were set to ρ1 = 50λ/ max(x) and ρ2 = ρ1/100 by empirical trial and error. A Gaussian noise model was used to reconstruct noise-free simulations while the Pois- son model was used for simulations incorporating shot- noise. All reconstructions (both Hawk and ADMM) were run for 1500 global iterations using HIO update with β = 0.9. For 500 additional iterations, the out-of-support update was switched from HIO to ER. The x-update ADMM step ran for 20 internal iterations in each global ADMM iteration. Reconstructions were rotated to ap- pear "upright" in the case that the algorithm produced an upside-down image, which is expected to occur in half of randomly seeded reconstructions due to the inversion (Friedel) symmetry of real diffraction images. Fig. 1a demonstrates the importance of support esti- mation. The low brightness of the camera man's black coat produces holes in the support estimated by Hawk, whereas our modified Shrinkwrap fills these in, leading to a better reconstruction. Fig. 1b, showing reconstruc- tions of the "Lena" from simulations incorporating mod- est shot-noise, demonstrates how the Poisson model and prior constraints jointly improve the final reconstruction. Fig. 1c shows how the error between the reconstruction and the ground truth improves during algorithm itera- tions. Our method converges in a comparable number of iterations to Hawk, but typically finds lower-error solu- tions. For a more experimentally relevant test, we simu- lated diffraction from a caffeine molecule's electron den- sity [27]. Fig. 2 shows the reconstructed results from measurements at different shot-noise levels. As shot noise increases, Hawk increasingly suffers from artifacts, where our method faithfully preserves the shape of molecule even with 10 times fewer photons. Finally, we assess the proposed algorithm on exper- imentally measured mimivirus diffraction patterns ob- tained at LCLS [4, 28]. The 512 × 512 px measurement contains a missing sphere of radius 35 px and a 20 px tall horizontal missing slit across the center. The RAAR al- gorithm was applied to allow comparison with published results [4]. We employed the area-based Shrinkwrap al- gorithm and RAAR parameters used in that work. Fig. 3 shows the reconstructed mimivirus from simu- lations at different shot-noise levels, where we modeled the original measurement as noise-free ground-truth and 4 (a) (b) (cid:21)(cid:23)(cid:21)(cid:20)(cid:15)(cid:18) (cid:17) t l u s e R M M D A t l u s e R k w a H 40.5 photons/px 4.05 photons/px 0.90 photons/px 0.405 photons/px (c) FIG. 3. (a) Experimentally recorded far-field diffraction pat- tern of a single mimivirus. (b) Reconstructed mimivirus from the original measurement using Hawk's implementation of RAAR. (c) Reconstructed mimivirus from shot-noise simu- lations at average [40.5, 4.05, 0.90, 0.405] photons/px, for our method (ADMM) and the Hawk package using σinit = 10, σend = 1. sampled it with Poisson statistics to mimic the diffrac- tion from a smaller object or weaker source. Employing Hawk's RAAR implementation, a drop in reconstruction quality occurs at 0.90 photons/px, and the virus becomes completely irretrievable at 0.405 photons/px. ADMM RAAR preserves the approximate shape of virus at 0.405 photons/px, with one missing and one attenuated edge of the hexagon-like shape of the projected virus. In conclusion, we propose a phase retrieval framework that allows simultaneous optimization of both the pri- mary model (here, Poisson or Gaussian pixel noise) and the prior constraints, with proper handling of missing measurements. It is efficiently solved using ADMM, by splitting the objective into sub-problems and addressing them independently. The decoupled treatment and the flexibility of adding/dropping priors at run time provides a significant productivity advantage. Together with a modified Shrinkwrap support update, the proposed algo- rithm produces high-quality reconstructions, in particu- lar for CXDI measurements with low photon counts. This work was funded through the LCLS Directorate of SLAC National Accelerator Laboratory under DoE BES Contract No. DE-AC02-76SF00515 (LS and TL) and an NSF CAREER Award No. IIS 155333 (GW). Thanks to Daniel Ratner and Filipe Maia for comments on a draft. TL would like to acknowledge Sara Salha and Kevin Raines for sharing thoughts on Bayesian phase re- trieval (as in eq. 1, manuscript forthcoming). ∗ [email protected] [1] J. Miao, P. Charalambous, J. Kirz, and D. Sayre, Nature 400, 342 (1999). [2] S. Marchesini, H. He, H. N. Chapman, S. P. Hau-Riege, and J. C. H. A. Noy, M. R. Howells, U. Weierstall, Spence, Phys. Rev. B 68, 140101 (2003). [3] P. Thibault, V. Elser, C. Jacobsen, D. Shapiro, and D. Sayre, Acta Crystallographica Section A: Foundations of Crystallography 62, 248 (2006). [14] Y. Nishino, J. Miao, and T. Ishikawa, 5 [4] M. M. Seibert, T. Ekeberg, F. R. N. C. Maia, M. Svenda, J. Andreasson, O. Jonsson, D. Odi´c, B. Iwan, A. Rocker, D. Westphal, M. Hantke, D. P. DePonte, A. Barty, J. Schulz, L. Gumprecht, N. Coppola, A. Aquila, M. Liang, T. A. White, A. Martin, C. Caleman, S. Stern, C. Abergel, V. Seltzer, J.-M. Claverie, C. Bostedt, J. D. Bozek, S. Boutet, A. A. Miahnahri, M. Messerschmidt, J. Krzywinski, G. Williams, K. O. Hodgson, M. J. Bo- gan, C. Y. Hampton, R. G. Sierra, D. Starodub, I. An- dersson, S. Bajt, M. Barthelmess, J. C. H. Spence, P. Fromme, U. Weierstall, R. Kirian, M. Hunter, R. B. Doak, S. Marchesini, S. P. Hau-Riege, M. Frank, R. L. Shoeman, L. Lomb, S. W. Epp, R. Hartmann, D. Rolles, A. Rudenko, C. Schmidt, L. Foucar, N. Kimmel, P. Holl, B. Rudek, B. Erk, A. Homke, C. Reich, D. Pietschner, G. Weidenspointner, L. Struder, G. Hauser, H. Gorke, J. Ullrich, I. Schlichting, S. Herrmann, G. Schaller, F. Schopper, H. Soltau, K.-U. Kuhnel, R. Andritschke, C.-D. Schroter, F. Krasniqi, M. Bott, S. Schorb, D. Rupp, M. Adolph, T. Gorkhover, H. Hirsemann, G. Potdevin, H. Graafsma, B. Nilsson, H. N. Chapman, and J. Hajdu, Nature 470, 78 (2011). [5] R. W. Gerchberg and W. O. Saxton, Optik (Stuttgart) Phys. Rev. B 68, 220101 (2003). [15] J. Miao, Y. Nishino, Y. Kohmura, B. son, C. Song, S. H. Risbud, Phys. Rev. Lett. 95, 085503 (2005). and T. John- Ishikawa, [16] X. Huang, J. Nelson, J. Steinbrener, J. Kirz, J. J. Turner, and C. Jacobsen, Opt. Express 18, 26441 (2010). [17] S. Salha, Inference from Incomplete Data in Coher- ent Diffraction Imaging, Ph.D. thesis, UCLA (2014), http://www.escholarship.org/uc/item/75x1988b. [18] K. He, M. K. Sharma, Opt. Express 23, 30904 (2015). and O. Cossairt, [19] L. I. Rudin, S. Osher, and E. Fatemi, Physica D: Non- linear Phenomena 60, 259 (1992). [20] S. Boyd, N. Parikh, E. Chu, B. Peleato, and J. Eck- stein, Foundations and Trends R(cid:13) in Machine Learning 3, 1 (2011). [21] P. Netrapalli, P. Jain, and S. Sanghavi, Signal Process- ing, IEEE Transactions on 63, 4814 (2015). [22] D. S. Weller, A. Pnueli, G. Divon, O. Radzyner, Y. C. Eldar, and J. A. Fessler, Computational Imaging, IEEE Transactions on 1, 247 (2015). [23] J. R. Fienup, Applied optics 32, 1737 (1993). [24] P. Soille, Morphological Image Analysis: Principles and 35 (1972). Applications. [6] J. R. Fienup, Appl. Opt. 21, 2758 (1982). [7] H. H. Bauschke, P. L. Combettes, J. Opt. Soc. Am. A 20, 1025 (2003). and D. R. Luke, [8] D. R. Luke, Inverse Problems 21, 37 (2005). [9] S. Marchesini, Review of scientific instruments 78, 011301 (2007). [10] D. Sayre, Acta Crystallographica 5, 843 (1952). [11] S. Ikeda and H. Kono, Opt. Express 20, 3375 (2012). [12] P. Thibault and M. Guizar-Sicairos, New Journal of Physics 14, 063004 (2012). [25] N. Parikh and S. P. Boyd, Foundations and Trends in optimization 1, 127 (2014). [26] F. R. Maia, T. Ekeberg, D. Van Der Spoel, and J. Hajdu, Journal of applied crystallography 43, 1535 (2010). [27] H. M. Berman, T. Battistuz, T. N. Bhat, W. F. Bluhm, P. E. Bourne, K. Burkhardt, Z. Feng, G. L. Gilliland, L. Iype, S. Jain, P. Fagan, J. Marvin, D. Padilla, V. Ravichandran, B. Schneider, N. Thanki, H. Weissig, J. D. Westbrook, and C. Zardecki, Acta Crystallograph- ica Section D: Biological Crystallography 58, 899 (2002). [13] R. G. Paxman, T. J. Schulz, and J. R. Fienup, [28] F. R. Maia, Nature methods 9, 854 (2012). J. Opt. Soc. Am. A 9, 1072 (1992).
1112.3492
2
1112
2012-01-02T13:24:45
Thin Films of Chiral Motors
[ "physics.bio-ph", "cond-mat.soft" ]
Hydrodynamic flows in biological systems are often generated by active chiral processes near or on surfaces. Important examples are beating cilia, force generation in actomyosin networks, and motile bacteria interacting with surfaces. Here we develop a coarse grained description of active chiral films that captures generic features of flow and rotation patterns driven by chiral motors. We discuss force and torque balances within the film and on the surface and highlight the role of the intrinsic rotation field. We arrive at a two dimensional effective theory and discuss our results in the context of ciliary carpets and thin films of bacterial suspensions.
physics.bio-ph
physics
Thin Films of Chiral Motors M. Strempel1,2, S. Furthauer1,2, S. W. Grill1,2, F. Julicher1 1Max Planck Institute for the Physics of Complex Systems, Nothnitzer Strasse 38, 01187 Dresden, Germany and 2Max Planck Institute of Molecular Cell Biology and Genetics, Pfotenhauerstr. 108, 01307 Dresden, Germany Abstract Hydrodynamic flows in biological systems are often generated by active chiral processes near or on surfaces. Important examples are beating cilia, force generation in actomyosin networks, and motile bacteria interacting with surfaces. Here we develop a coarse grained description of active chiral films that captures generic features of flow and rotation patterns driven by chiral motors. We discuss force and torque balances within the film and on the surface and highlight the role of the intrinsic rotation field. We arrive at a two dimensional effective theory and discuss our results in the context of ciliary carpets and thin films of bacterial suspensions. 2 1 0 2 n a J 2 ] h p - o i b . s c i s y h p [ 2 v 2 9 4 3 . 2 1 1 1 : v i X r a 1 The building blocks of biological systems are chiral molecules such as proteins and DNA. The emergent collective behaviors of these molecular components give rise to active dynamic processes in cells and tissues which can also reflect these molecular chiralities. A key example is the breaking of left-right symmetry in organisms with a well-defined handedness during development [1 -- 3] (i.e. the heart is on the left side in humans). Many pattern forming events in biological systems are governed by active chiral processes. For example, it has been shown that the rotating beat of cilia, which drives chiral hydrodynamic flows, is at the basis of left-right symmetry breaking in vertebrate animals [4 -- 6]. In addition, active chiral processes have been observed in the cytoskeleton of cells [7, 8]. The cytoskeleton is a gel-like network of elastic chiral filaments and other components such as motor proteins. The interactions of motors and filaments drive movements and intracellular flows [9, 10], which can have chiral asymmetries [7, 11 -- 13]. Active chiral processes are typically observed on surfaces or at interfaces. Key examples are carpets of beating cilia driving hydrodynamic flows parallel to the surface, on which they are attached [14, 15], and rotating motors on a surface [16 -- 18]. Many different beating patterns of cilia exist, which in general are chiral and often exhibit rotating movements, as is the case for the cilia that govern the left-right symmetry breaking of organisms [6, 19]. Many microorganisms possess carpets of cilia on their outer surface, which are used for self-propulsion along helical trajectories in a fluid [15]. Recently, the collective behavior of swimming E. coli bacteria on solid surfaces was studied [20]. These bacteria possess rotary motors, which drive the rotation of helical flagella relative to the cell body. This relative rotation provides a motor for self-propulsion in a fluid. Due to conservation of angular momentum, cell body and flagella exert equal but opposite torques on the surrounding fluid, see Fig. 1(a). Thus, the bacteria act as torque dipoles. Such flagellated E. coli were placed on a solid agar surface [20], where they produced a thin fluid layer that promotes their motility. Some bacteria attach to the surface while others swim in the film. Those bacteria which were attached to the surface also exerted torques on the solid substrate. Large scale chiral flow patterns were reported as a result of bacterial activity [20]. The bacterial suspension therefore represents an internally driven active chiral film supported by a surface. Active fluids and gels have been studied in the framework of hydrodynamic theories both as a paradigm for the cell cytoskeleton and for suspensions of active swimmers [21 -- 26]. 2 Such approaches are based on liquid crystal hydrodynamics [27 -- 32] driven out equilibrium by internal active processes. It has been shown that stresses generated by active processes can give rise to a rich variety of dynamic patterns and flows [10, 33 -- 37]. Similar approaches have also been used for the study of granular systems [38, 39]. Chiral effects have so far been discussed mainly in passive systems and in granular gases [40]. A study of active chiral processes in fluids and gels is so far lacking. Here, we develop a generic theory for active chiral films. We consider a thin film of a suspension of chiral motors on a solid surface. Our work is based on a general discussion of the bulk properties of active chiral fluids [41]. Active chiral processes are introduced as torque dipoles arising from counterrotating objects, such as the cell body and the flagella of bacteria, see Fig. 1(a). We start by discussing the bulk properties of a fluid with mass density ρ and center of mass velocity v in which active processes take place. Linear momentum and angular momentum conservation can be expressed as ∂t(ρvα) = ∂βσtot αβ + f ext α , ∂tltot αβ = ∂γM tot αβγ + τ ext αβ + rαf ext β − rβf ext α , (1) (2) where Einstein's summation convention is implied. Here ρvα is the momentum density. The density of angular momentum ltot αβ is described by an antisymmetric second rank tensor. Externally applied force and torque densities are denoted by f ext α and τ ext αβ , respectively, rα is a position vector, σtot αβ is the total stress and M tot αβγ is the total angular momentum flux. The total stress σtot αβ = −P δαβ + σs αβ + σa αβ describes momentum fluxes and can be decomposed in the isotropic pressure P , the symmetric traceless stress σs αβ and the antisymmetric stress σa αβ. The total angular momentum density consists of an orbital part ρ(rαvβ − rβvα) and a spin angular momentum density lαβ = ltot αβ − ρ(rαvβ − rβvα), see [27, 28, 32, 40]. Similarly, the total angular momentum flux M tot αβγ = Mαβγ + (rασtot βγ − rβσtot αγ ) is the sum of fluxes Mαβγ and rασtot βγ − rβσtot αγ of spin and orbital angular momentum, respectively. Note that the spin angular momentum density lαβ and the spin angular momentum flux Mαβγ do not depend on the choice of coordinate system. We define an effective rate of intrinsic rotation Ωαβ of local volume elements via lαβ = IαβγδΩγδ, where Iαβγδ is the moment of inertia tensor per unit volume [32]. The balance of the local torque then reads ∂tlαβ = −2σa αβ + ∂γMαβγ + τ ext αβ . (3) 3 FIG. 1: (a) Schematic representation of a swimming bacterium with rotating flagella. (b) Chiral motor consisting of two counterrotating spheres at distance d. Arrows indicate senses of rotation. Eq. (3) shows that spin angular momentum is not conserved but can be exchanged with orbital angular momentum by antisymmetric stress. Active chiral processes introduce contributions to the total stress and to fluxes of spin angular momentum. We discuss the effect of an active chiral process by considering a torque dipole ταβ(r) in a fluid, located at r = 0, see Fig. 1(b). This torque dipole is built from two torque monopoles ±qǫαβγpγ, of strength q separated by a small distance d in the direction of the torque axis given by the unit vector p: ταβ = qǫαβγpγ (cid:18)δ(r − d 2 p) − δ(r + d 2 p)(cid:19) ≃ −qdǫαβνpνpγ∂γδ(r) . (4) Note that ταβ is invariant under the transformation p → −p, which implies a nematic character. Active internal torques and forces are introduced similarly to external ones. Therefore the torque dipole ταβ can be interpreted as an active contribution M act αβγ = −qdǫαβδpδpγδ(r) to the spin angular momentum flux, which obeys ∂γM act αβγ = ταβ. In a suspension of many identical torque dipoles at positions r(i) and with orientations p(i) the active angular momentum fluxes are M act αβγ = −qdXi ǫαβδp(i) δ p(i) γ δ(r − r(i)) 4 ≃ ζǫαβδpδpγ + ζ ′ǫαβγ , (5) where ζ = −Sqdn(r) and ζ ′ = (S − 1)qdn(r)/3 describe the strength of a nematic and an isotropic active contribution to Mαβγ, respectively, and n(r) is the dipole density. Here and below p, with p2 = 1, denotes a coarse grained nematic director and S is a nematic order parameter which obeys S(pαpβ − (1/3)δαβ) =< p(i) β > −(1/3)δαβ, where the average is α p(i) taken over a volume element [30]. The constitutive equation for the spin angular momentum flux then has the form Mαβγ = κ∂γΩαβ + ζǫαβδpδpγ + ζ ′ǫαβγ . (6) Here the term proportional to the phenomenological coefficient κ describes passive dissipative processes. Furthermore we ignore other passive couplings [32, 41]. Here and below we have neglected inertial terms. We complement Eq. (6) by the constitutive equations for the stresses of a passive incompressible fluid: σs αβ = 2ηuαβ , σa αβ = 2η ′(Ωαβ − ωαβ) , (7) (8) where uαβ is the traceless part of the strain rate uαβ = (∂αvβ + ∂βvα) /2 and ωαβ = (∂αvβ − ∂βvα) /2. Here, η is the shear viscosity and η ′ is a rotational viscosity. The pressure P plays the role of a Lagrange multiplier and is imposed by the incompressibility condition ∂γvγ = 0. Other coupling terms, known to exist in liquid crystals, have been neglected for simplicity. We now study the stresses and flows generated by active chiral processes in a thin fluid film of height h on a solid substrate, see Fig. 2. Note that the film is not symmetric with respect to z → h − z because the two surfaces at z = 0 and z = h differ. Integrating the force balance Eq. (1) we obtain 0 = 1 h h Z 0 dz∂βσtot iβ = ∂j ¯σtot ij + 1 h σtot iz h 0 , (9) iβ ≡ (1/h)R h 0 dzσtot iβ , and σtot iz h where the bar denotes an average over the film height, ¯σtot 0 ≡ iz (z = 0), where i = x, y and j = x, y, see Fig. 2. The shear stress at the σtot iz (z = h) − σtot film surfaces is written as 1 h σtot iz h 0 = −ξt¯vi + ξΩ ¯Ωiz , 5 (10) which describes friction forces due to relative motion with the substrate by the coefficient ξt and due to relative rotation with the substrate by the coefficient ξΩ. From the torque balance Eq. (3) and ignoring inertial terms, we find 0 = −2¯σa αβ + ∂j ¯Mαβj + 1 h Mαβzh 0 . The torque densities on the surfaces are 1 h Mαβzh 0 = −ξr ¯Ωαβ + ζsǫαβδ ¯pz ¯pδ + ζ ′ sǫαβz (11) (12) , where frictional torques are described by the coefficient ξr. The torque exerted by active chiral processes on the boundaries is described by the coefficients ζs and ζ ′ s. Averaging over the height of the film also leads to new coefficients for the active in-film angular mo- mentum fluxes ¯Mαβj = ζf ǫαβδ ¯pδ ¯pj + ζ ′ the boundaries is σext f ǫαβj. Finally, the externally applied shear stress on In the absence of external shear 0 = −2hη ′ ¯Ωiz/(η + η ′). Similarly, ¯σzz = P ext is the externally applied pressure stresses vih and therefore ¯P = −2η∂i¯vi. We obtain equations for the flow and rotation field, iz ≈ ¯σiz = (η + η ′)vih 0/h + 2η ′ ¯Ωiz. 0 = ∂j [(3η − η ′)∂i¯vj + (η + η ′)∂j ¯vi] + 2η ′∂j ¯Ωij − ξt¯vi + ξΩ ¯Ωiz , 0 = −(4η ′ + ξr) ¯Ωxy + 2η ′(∂x¯vy − ∂y ¯vx) + κ∂2 0 = −(cid:18) 4ηη ′ η + η ′ + ξr(cid:19) ¯Ωiz + κ∂2 j j ¯Ωiz + ∂jζf ǫizδ ¯pδ ¯pj + ∂jζ ′ ¯Ωxy + ∂jζf ǫxyz ¯pz ¯pj + ζsǫxyz ¯p2 z + ζ ′ sǫxyz f ǫizj + ζsǫizδ ¯pδ ¯pz , (13) ,(14) (15) where we used the thin film approximation ∂ivz ≪ ∂zvi. Motivated by rotating cilia or bacteria attached to a surface, we now consider the case of a thin active film containing chiral motors that are on average aligned along the vector ¯p = (cos φ sin θ, sin φ sin θ, cos θ). Here we have introduced the angle θ which describes the average tilt with respect to the surface normal vector and the angle φ which specifies the tilt direction with respect to the x-axis. We first consider a homogeneous distribution of motors in an infinite system where all spatial derivatives vanish. In this case Eq. (13) reads ¯vi = (ξΩ/ξt) ¯Ωiz. Using Eq. (15) we find ¯vi = ξΩζs ξt ǫizj ¯pj ¯pz ξr + 4ηη ′/(η + η ′) . (16) A flow with a velocity ¯v ∝ sin(2θ) is generated. The flow direction is perpendicular to the direction of tilt with respect to the surface normal vector. This case describes the collective 6 FIG. 2: Schematics of a thin fluid film of height h that contains chiral motors. Motors are torque dipoles that consist of counterrotating spheres (see Fig. 1(b)), one of which is attached to the surface. The other rotates as indicated by the arrows. The average motor direction is described by the vector ¯p which is tilted at an angle θ in y-direction. generation of flow by carpets of cilia on a surface [5, 6, 19]. Note that ξΩ and therefore also the generated flow vanish in films that are symmetric with respect to z → h − z. Ciliary carpets do have this asymmetry and can therefore generate flows [6]. Furthermore, we find from Eq. (14) an intrinsic rotation rate ¯Ωxy = (ζs cos2 θ + ζ ′ s)/(4η ′ + ξr) generated by the rotating motors with vorticity ¯ωxy = 0 in the hydrodynamic flow field. As a second example we consider a circular patch of radius R that contains active motors. This is described by position dependent coefficients ζs(r) = ζsΘ(R − r) and by similar expressions for ζ ′ s, ζf and ζ ′ f , where Θ(r) is the Heaviside function. We numerically solve Eqs. (13), (14) and (15) with periodic boundary conditions for boxsize L = 4R using Fourier transforms, see Fig. 3. The velocity field for a tilt angle θ = 0 is displayed in Fig. 3(a). The patch generates a chiral flow field driven by the intrinsic rotations ¯Ωxy. The stream lines are concentric circles. The flow velocity is maximal at the edge of the patch and decays exponentially outside. There is no net transport across the patch. If the motors are tilted with θ 6= 0 along the y axis, a net transport in x direction across the patch with velocity proportional to sin(2θ) appears in addition to the circular flow, see Fig. 3(b). We have developed a coarse grained description of flow patterns generated in thin active films in which chiral processes take place. We have shown that if the chiral motors are tilted with respect to the surface normal vector, directed flows can be generated over large 7 FIG. 3: Flow fields (vectors) generated by a circular patch of chiral motors (blue). a) Average motor axis ¯p is perpendicular to the surface, i.e. θ = 0. b) Average motor axis ¯p tilted at angle θ = π/4 in y-direction. Flow fields were determined numerically in a square box of size L with periodic boundary conditions. Parameter values are η′ = η = h2κ = h2ξt = −hξΩ = ξr, ¯ζ = −hζs = 10¯ζ ′ = −10hζ ′ s, L = 20h. distances in a direction perpendicular to the tilt direction. The velocity of this net flow is maximal for tilt angles of θ = 45◦. This result accounts for the generation of net flows and left-right symmetry breaking by carpets of cilia in the mouse ventral node [4 -- 6] and Kupffer's vesicle in zebrafish [42]. Note that in both cases a tilt angle with respect to the surface normal can be defined [19]. Interestingly, reported tilt angles vary between 30◦ and 50◦ [43], which is close to the angle of maximum transport velocity. Moreover in the case of finite patches of chiral motors, intrinsic rotation rates drive chiral flows along the edge of the patch. This phenomenon was reported in recent experiments on bacterial films on solid surfaces [20]. Our theory highlights the role of the intrinsic rotation field Ωαβ for active chiral pro- cesses. In particular we find net flows without vorticity generated by intrinsic rotations in a homogeneous system, see Eq. (16). Note that in passive bulk fluids Ωαβ converges to the flow vorticity after a relaxation time that in general is short [32]. In films of chiral motors, however, the intrinsic rotation rate differs from the vorticity even at steady state, and can create effects near interfaces and surfaces. Our work shows that the hydrodynamic flow velocity v and the intrinsic rotation rate Ωαβ in active chiral films are coupled by the ac- tive processes and by the boundary conditions. This can give rise to complex flow patterns 8 generated by carpets of chiral motors. Our theory permits a simplified description of the flows generated by ciliary carpets. It will be interesting to expand this theory to include phase relationships between adjacent motors to capture the self-organization of beating cilia in metachronal waves [15]. [1] L. Wolpert, Principles of development (Oxford University Press, Oxford, 1998). [2] C. L. Henley, Possible mechanisms for initiating macroscopic left-right asymmetry in devel- oping organisms, arXiv:0811.0055v2 (2008) [3] L. N. Vandenberg and L. Levin, Perspectives and open problems in the early phases of leftright patterning, Seminars in Cell and Developmental Biology 20, 456-463 (2009). [4] S. Nonaka, Y. Tanaka, Y. Okada, S. Takeda, A. Harada, Y. Kanai, M. Kido, and N. Hirokawa, Randomization of left-right asymmetry due to loss of nodal cilia generating leftward ow of extraembryonic uid in mice lacking KIF3B motor protein, Cell 95, 829-837 (1998). [5] S. Nonaka, S. Yoshiba, D. Watanabe, S. Ikeuchi, T. Goto, W.F. Marshall, H. Hamada, De Novo Formation of Left/Right Asymmetry by Posterior Tilt of Nodal Cilia, PLoS Biol 3 , e268 (2005). [6] J. Buceta, M. Ibanes, D. Rasskin-Gutman, Y. Okada, N.Hirokawa, J. C. Izpis´ua-Belmontey, Nodal cilia dynamics and the specification of the left/right axis in early vertebrate embryo development, Biophysical Journal 89, 2199-2209 (2005). [7] M. Danilchik, E. E. Brown, K. Riegert, Intrinsic chiral properties of the Xenopus egg cortex: an early indicator of left-right asymmetry?, Development 133, 4517-4526 (2006). [8] J. Xu, A. Van Keymeulen, N. M. Wakida, P. Carlton, M. W. Berns, H. R. Bourne, Polarity reveals intrinsic cell chirality, Proc. Natl. Acad. Sci. USA 22, 9296-9300 (2007). [9] J. W. van de Meent, I. Tuval, R. E. Goldstein, Natures Microfluidic Transporter: Rotational Cytoplasmic Streaming at High P´eclet Numbers, Phys. Rev. Lett. 101, 178102 (2008). [10] M. Mayer, M. Depken, J. S. Bois, F. Julicher, S. W. Grill, Anisotropies in cortical tension reveal the physical basis of polarizing cortical flows, Nature 467, 617-621 (2010). [11] I. Sase, H. Miyati, S. Ishiwata, and K. Kinosita Jr., Axial rotation of sliding actin filaments revealed by single-fluorophore imaging, Proc. Natl . Acad. Sci . USA 94, 5646-5650 (1997). [12] A. Hilfinger and F. Julicher, The chirality of ciliary beats, Phys. Biol. 5, 016003 (2008). 9 [13] A. Vilfan, Twirling motion of actin filaments in gliding assays with nonprocessive Myosin motors, Biophysical Journal 97, 1130-1137, (2009). [14] I. R. Gibbons, Cilia and flagella of eukaryotes, J. Cell Biol. 91, 107s-124s (1981). [15] D. Bray, Cell movements (Garland Publishing, New York, 2001), 2nd ed. [16] H. Noji, R. Yasuda, M. Yoshida, K. Kinosita Jr, Direct observation of the rotation of F1- ATPase, Nature 386, 299-302 (1997). [17] P. Lenz, J. F. Joanny, F. Julicher, J. Prost, Membranes with Rotating Motors, Phys. Rev. Lett. 91, 108104 (2003). [18] N. Uchida and R. Golestanian, Synchronization and Collective Dynamics in A Carpet of Microfluidic Rotors, Phys. Rev. Lett. 104, 178103 (2010). [19] D. J. Smith, J. R. Blake, E. A. Gaffney, Fluid mechanics of nodal flow due to embryonic primary cilia, J. R. Soc. Interface 5, 567-573 (2008). [20] Y. Wu, B. G. Hosu, H. C. Berg, Microbubbles reveal chiral fluid flows in bacterial swarms, Proc. Natl. Acad. Sci. USA 10, 4147-4151 (2011). [21] R. A. Simha, S. Ramaswamy, Hydrodynamic Fluctuations and Instabilities in Ordered Sus- pensions of Self-Propelled Particles, Phys. Rev. Lett. 89, 058101 (2002). [22] T. B. Liverpool and M. C. Marchetti, Instabilities of Isotropic Solutions of Active Polar Filaments, Phys. Rev. Lett. 90, 138102 (2003). [23] I. S. Aranson and L. S. Tsimring, Model of coarsening and vortex formation in vibrated gran- ular rods, Phys. Rev. E 67, 021305 (2003). [24] Y. Hatwalne, S. Ramaswamy, M. Rao, R. A. Simha, Rheology of active-particle suspensions, Phys. Rev. Lett. 92, 118101 (2004). [25] K. Kruse, J. F. Joanny, F. Julicher, J. Prost and K. Sekimoto, Asters, Vortices, and Rotating Spirals in Active Gels of Polar Filaments, Phys. Rev. Lett. 92, 078101 (2004). [26] K. Kruse, J. F. Joanny, F. Julicher, J. Prost, and K. Sekimoto, Generic theory of active polar gels: a paradigm for cytoskeletal dynamics, Eur. Phys. J. E 16, 5 (2005). [27] J. L. Ericksen, Anisotropic Fluids, Arch. Ration. Mech. Anal. 4, 231 (1960). [28] F. M. Leslie, Some constitutive equations for anisotropic fluids, Q. J. Mech. Appl. Math. 19, 357 (1966). [29] P. C. Martin, O. Parodi, P. S. Pershan, Unified Hydrodynamic Theory for Crystals, Liquid Crystals, and Normal Fluids, Phys. Rev. A. 6, 6 (1972). 10 [30] P. G. de Gennes and J. Prost, The Physics of Liquid Crystals, Vol. 2, (Oxford University Press, Oxford, 1995). [31] H. Stark and T. C. Lubensky, Poisson-bracket approach to the dynamics of nematic liquid crystals, Phys. Rev. E 67, 061709 (2003). [32] H. Stark and T. C. Lubensky, Poisson-bracket approach to the dynamics of nematic liquid crystals: the role of spin angular momentum, Phys. Rev. E 72, 051714 (2005). [33] R. Voituriez, J. F. Joanny, J. Prost, Spontaneous flow transition in active polar gels, Europhys. Lett. 96, 028102 (2005). [34] G. Salbreux, J. Prost, J. F. Joanny, Hydrodynamics of Cellular Cortical Flows and the For- mation of Contractile Rings, Phys. Rev. Lett. 103, 058102 (2009). [35] J. S. Bois, F. Julicher, S. W. Grill, Pattern formation in active fluids, Phys. Rev. Lett. 106, 028103 (2011). [36] L. Giomi, L. Mahadevan, B. Chakraborty, M. F. Hagan, Excitable Patterns in Active Nematics, Phys. Rev. Lett 106, 218101 (2011). [37] J. Elgeti, M. E. Cates, D. Marenduzzo, Defect hydrodynamics in 2D polar active fluids, Soft Matter 7, 3177 (2011). [38] L. Bocquet, W. Losert, D. Schalk, T. C. Lubensky, J. P. Gollub, Granular shear flow dynamics and forces: Experiment and continuum theory, Phys. Rev. E 65, 011307 (2001). [39] I. S. Aranson and L. S. Tsimring, Theory of self-assembly of microtubules and motors, Rev. Mod. Phys. 78, 641-692 (2006). [40] J. C. Tsai, F. Ye, J. Rodriguez, J. P. Gollub, T. C. Lubensky, A Chiral Granular Gas, Phys. Rev. Lett. 94, 214301 (2005). [41] S. Furthauer, M. Strempel, S. W. Grill, F. Julicher, Generic theory of Active chiral fluids (in preparation). [42] J. J. Essner, J. D. Amack, M. K. Nyholm, E. B. Harris, H. J. Yost, Kupffer's vesicle is a ciliated organ of asymmetry in the zebrafish embryo that initiates left-right development of the brain, heart and gut, Development 132, 1247-1260 (2005). [43] Y. Okada, S. Takeda, Y. Tanaka, J. C. Izpis´ua Belmonte, N. Hirokawa, Mechanism of Nodal Flow: A Conserved Symmetry Breaking Event in Left-Right Axis Determination, Cell 121, 633-644 (2005). 11
1808.00108
2
1808
2018-10-26T08:03:46
The growth and development of living organisms from the thermodynamic point of view
[ "physics.bio-ph" ]
The living organism is considered as an open system, whereas Prigogine's approach to the thermodynamics of such systems is used. The approach allows one to formulate the law of individual growth and development (ontogenesis) of the living organism, whereas it has taken into account that the development and functioning of the system are occurring under the special internal program. The thermodynamic equation of growth is followed a method of estimation of the specific entropy of organism. The theory is compared with experimental data, whereas estimates of the specific entropy of some species were done; it shows a reduction of specific entropy in the evolutionary row: yeast - insects - reptiles - birds.
physics.bio-ph
physics
The growth and development of living organisms from the thermodynamic point of view Alexei A. Zotin Koltzov' Institute of Developmental Biology RAS Vavilov Str. 26, Moscow, Russia 119334 and Vladimir N. Pokrovskii Moscow State University of Economics, Statistics and Informatics, Moscow, RUSSIA 119501 Abstract The living organism is considered as an open system, whereas Pri- gogine's approach to the thermodynamics of such systems is used. The approach allows one to formulate the law of individual growth and de- velopment (ontogenesis) of the living organism, whereas it has taken into account that the development and functioning of the system are occurring under the special internal program. The thermodynamic equation of growth is followed a method of estimation of the specific entropy of organism. The theory is compared with experimental data, whereas estimates of the specific entropy of some species were done; it shows a reduction of specific entropy in the evolutionary row: yeast - insects - reptiles - birds. Keywords: Entropy of living organism, Energy exchange during growth, Thermodynamic equation of growth October 2018 1 1 Introduction Many prominent investigators [1 - 4] approached the formulation of phe- nomenological (thermodynamic) theory of the development, growth and ag- ing of a living organism (theory of ontogenesis). The theories were based on the analysis of quantities characterizing energy exchange between an or- ganism and the surrounding. Measuring simultaneously both the incoming flux of energy and intensity of heat production, the investigators found [5, 6] that, under the growth, the former can surpass the later. The difference be- tween the incoming flux of energy and intensity of heat production, dubbed as ψu-function [5, 6], has been qualified as a part of energy that is needed for the construction and growth of the organism [7]. This idea was central in our investigation, and we are going now to formulate the exact relation between ψu-function and the rate of mass growth of the living organism. From the thermodynamic point of view, the living organism ought to be considered as an open thermodynamic system that exists in a non- equilibrium state [4, 8 - 10], and the desire to understand the general princi- ples that govern the living organisms was one of the important motivations to investigate open systems. The living organism does not contain anything, apart of chemical compounds, and all processes that are occurring in the sys- tem are connected, in the large degree, with chemical transformations, so that there were some attempts to present simple models imitating the be- haviour of the living organism. The brusselator, considered by Prigogine with collaborators [11], can be considered as one of the elementary models that reflects the essential features of the biological organism: an exchange of substances with the environment and existence of stable spatial internal structure within a limited volume. However, the living organism has extra peculiarities of the behaviour: unlike the brusselator, all processes are run- ning under internal control, trying to keep the internal state of organism unchangeable, which is known as a phenomenon of homeostasis [12]. The control mechanism changes the speed of running of processes depending on the signals receiving from the outside. At the rise of outer temperature, for example, the control mechanism commands to increase perspiration, for the temperature of the organism not to be increased. This special congenital program also determines the development and growth of the organism. Do all of this require some modification of thermodynamic laws at appli- cation to living systems in comparison with chemical systems of the brusse- lator type? In this paper, first of all, we are going to remind main results of Prigogene's formulation of thermodynamics of open systems (Section 2). In Section 3 we apply Prigogene's results to show how the control apparatus 2 can be introduced into the thermodynamic approach (Section 3.1); this gives a clue to the problem of the description of growth and development of the living organism. Then, we are formulating (Section 3.3) and testing (Sec- tion 3.4) the thermodynamic equation of growth of the living organism. The Discussion and Conclusions section contains a discussion of the resolution of the problem presented herein. 2 The Change of Entropy in Prigogine's Formula- tion The essential contribution to the thermodynamics of open systems was brought by Prigogine, when he and his collabourators investigated systems of chemically reacting substances. Stationary states of such systems exist due to exchange both particles and energy with the environment. In section 8 of the third chapter of his book [9], Prigogine has specified three contri- butions to the variation of entropy of an open system. The simple analysis [13, 14], reproduced in Appendix, demonstrates that the change of entropy S of the system at the given temperature T can be calculated according to the formula T dS = ∆Q −(cid:88) (cid:88) Ξj ∆ξj + ηα ∆Nα. (1) j α The first term on the right-hand side of the equation presents a stream of thermal energy into the system; it can be positive (the stream into the sys- tem) or negative (the stream out of the system). The last term depicts a stream of energy coming with the stream of particles of substances ∆Nα that can be positive (the stream into the system) or negative (the stream out of the system), ηα is connected with chemical potential of substance labeled α. The middle term on the right-hand side of equation (1) presents a change of the stored energy inside the system due to the existing of inter- nal variables ξi, so called thermodynamic forces Ξj are considered positive, so that this term gives a positive contribution into entropy at relaxation of internal variables. In the case of chemically reacting substances, which was investigated by Prigogine [9], the internal variables ξi appear to be mea- sures of incompleteness of chemical reactions, that are the measures of how much the considered system with chemical reactions is out of equilibrium; the thermodynamic forces Ξj are affinities of corresponding reactions. The theory was generalized [13, 14] to consider any deviation from the equilib- rium state as an internal variable, so that we can consider the set of internal variables ξj in equation (1) to be comprised of the quantities determing not 3 only degrees of completeness of all chemical reactions occurring in the sys- tem, but also a structure of the system, gradients of temperature, difference of concentrations of substances and so on. Note that equation (1) presents the balance of energy and is not restricted to any approximations. The sign of the middle term on the right-hand side of equation (1) is taken in such a way to ensure the positive production of entropy at the relaxation of internal variables ξj, but two processes are running simultane- ously. Under some external influences (the fluxes of heat and substances) the system is passing to a new state, while internal variables could emerge and entropy of the system could change. Due to internal processes that cannot be controlled from the outside, the internal variables, as measures of non-equilibrium of the system, tend to disappear. For small deviations of internal variables ξi from their equilibrium values ξ(0) , the local law of disappearing can be written as relaxation equation for each internal variable i (cid:16) (cid:17) dξi dt = − 1 τi ξi − ξ(0) i , i = 1, 2, . . . , (2) where τi = τi(T, x1, x2, . . . , xn) is a relaxation time of a corresponding inter- nal variable that, being left on its own, permanently decreases and, eventu- ally, comes close to equilibrium value ξ0 i , which is conveniently considered to be equal to zero. In this situation, due to the recorded equation (2), the middle term on the right-hand side of equation (1), is positive and describes the dissipation of the stored energy. The dynamics of internal variables for arbitrary deviations from equilibrium was formulated in the paper [14]. Note that, in line with the notion of entropy of the open system S, one can introduce the notion of complexity; the latter has the same sense as notion of bound information [15] in application to complex systems. The complexity can be considered as a characteristic of the set of internal vari- ables ξi: the internal variables themselves sometimes are called variables of complexity or variables of order. Though at the moment it is impossible to define how complexity is related to the number and connectedness of internal variables, one can assert that the greater the number of internal variables, their values and mutual connections, the greater the complexity. In virtue of equation (1), the change of complexity can be connected with the change of entropy of the system: increase or decrease in number and values of internal variables provokes decrease or increase in entropy. 4 (cid:18) S (cid:19) M (cid:33) (cid:88) α (cid:32) −(cid:88) i (cid:88) α 3 Organism as an Open Thermodynamic System We shall discuss the processes of growth and development of the living organ- ism, considering it as an open thermodynamic system of undefined volume, but with the certain mass M and temperature T . To describe a huge number of internal processes within the system, one needs a set of internal variables ξj, that, in case of the living organism, are emerged under the influence of the streams of substances and heat. On the basis of fundamental expression for variation of entropy of an open system (1), the equation for the growth rate of specific entropy S/M of the organism can be written as M S d dt + 1 M dM dt = 1 T S ∆Q ∆t Ξi dξi dt + ηα ∆Nα ∆t . (3) It is marked with the symbol ∆ that the corresponding quantities are objects of direct empirical observation. The law of conservation of mass determines the equation of growth dM dt = mα ∆Nα ∆t , (4) where mα is mass of a particle of substance labeled α. Unfortunately, this form of the growth equation is useless practically, so as there are a lot of the processes occurring in the living system, and it is extremely difficult to measure the substances coming in and out the organism [4]. The balance equations (3) and (4) give a basis for thermodynamic inter- pretation of the phenomena of existence, growth and ageing of a biological organism, but for this purpose, it is necessary to establish, first of all, cor- respondence of all symbols in these parities to the quantities concerning the living organism. 3.1 Three groups of internal variables Let's remind, that the internal variables that are present in the middle term of the right part of the equation (3) determine the deviation of the system from the state of equilibrium and, in this way, apart of the incompleteness of internal chemical reactions, describe the architectural and functional struc- ture of the system. Considering the organism as an open system, we can note that there is permanent changing of internal variables, which emerge under fluxes of heat and substances and tend to disappear. The rate of disappearing is different for different internal variables; time of relaxation 5 of some of them are very small as compared with the lifespan of organism. The other internal variables exist during the total lifetime and even more. The difference of internal variables according to their rates of approach the equilibrium state determines their functional distinctions in the processes of the growth and development of the organism.1 For the further discussion, we are dividing the set of internal variables of the system into three groups, according to values of relaxation times; further on, the different functional roles of separate groups in the processes of development and growth are described. First of all, it is possible to separate a group of internal variables, which originates from the complexity of the first cell that gives the beginning of the organism. The material carriers of complexity are the molecules of a deoxyribonucleic acid packed in the nuclei of the cells. The copies of internal variables (complexity) of the initial set are present in each cell and remain constant (at some idealization) up to the death of an organism. The number of copies of these variables increases at the proliferation of cells, and during the realization of the program of development. These variables comprise the foundation of a control mechanism that supervises all processes in the organism during its existence and growth. The organism is developing under the installed plan received by right of succession, whereas some internal variables are suppressed or expressed when the cells are multiplying. A role of internal variables of the first group is the conservation of hereditary instructions and the control over the construction of an organism. The second group of internal variables describes the emerged structure of the organism in processes of morphogenesis and differentiation, when the interoperability of the parts of an organism is materializing. For example, conformations of the biomolecules having albuminous in the basis, change in such a manner that speed of reactions in which they participate, also sharply changes. The greater role is also playing a spatial arrangement of fibers and, in particular, enzymes [17]; it provides such processes that cannot occur in a random mixture of chemical reagents. The number and values of such internal variables increase at growth and development of the organism. The set of variables of the second groups determines some metastable structure (organs and tissues) of the living organism and have the times of relaxation comparable with the lifespan of the organism, so that they do not contribute into the current dissipation of energy (production of entropy). 1Consideration of an hierarchy of relaxation processes is useful in many applications of non-equilibrium thermodynamics. For example, Gujrati [16] demonstrated recently that the division of internal variables according to their relaxation times is very helpful for description of the phenomena of solidification of glasses. 6 The set of internal variables of the third group emerges in the course of the above-described processes of multiplication of original complexity and construction of the structure of the organism. The processes are based on chemical transformations of substances, occur in non-equilibrium situations and are accompanied by excitation of internal variables and their fast re- laxation. The set of internal variables of this group includes the variables locating gradients of temperature, differences in concentration of substances and degrees of completeness of all chemical reactions running in the organ- ism. Internal variables of the third group accompany all existing processes anywhere they were. These internal variables have very small times of re- laxation (in comparison with existence of the organism) and determine the basic contribution to the internal production of entropy. The dissipated energy is coming out in the form of heat. Note that now we can record an expression for entropy of the organism as a function of the internal variables, whereas the expansion can be done over internal variables of the third group ξ(3) S(ξ(1), ξ(2), ξ(3)) = S0(ξ(1), ξ(2)) − 1 2T Sij(ξ(2)) ξ(3) i ξ(3) j . The quantity S0 is entropy of a dead body (a corpse), the quantity is a decreasing function of variables of the first group ξ(1) (carriers of instructions for construction and functioning of an organism) and the second group ξ(2) (the description of metastable structures of the organism). The last term on the right hand side of the equation presents all processes running in the live organism; the quantities Sij are positive and essentially depend on internal variables of the second group, which define metastable structure (a set of structures) of the system and cannot be considered as small quantities. The difference in the relaxation times of internal variables allows us to distinguish different functional roles of internal variables in the processes of development and growth. The distribution of the all internal variables over three groups can be helpful for designing of the functional scheme of the living organism. 3.2 Exchange of energy between the organism and environ- ment There are methods of assessments of incoming and outgoing fluxes of energy for the organism of an animal [4]; these fluxes are presented in the right part of the equation (3) by the first and the last terms. The greatest amount of energy is delivered into organisms with food in the form of energy of chem- ical links of organic molecules [4, 18]. The speed of synthesis of adenosine 7 triphosphate (ATP) gives us an estimate of the coming-with-food energy (in a unit of time). The synthesis can be realized in the two basic processes: the oxidizing phosphorylation and glycolysis. One can assume that the energy received in the process of glycolysis, can be neglected in comparison with the process of oxidizing phosphorylation [4], and one can judge about the speed of synthesis ATP indirectly, namely on the rate of oxygen consump- tion or the rate of allocation of carbonic gas. It is shown, that quantity of the consumed oxygen is proportional to the energy coming into the organ- ism, whereas the factor of proportionality (so-called oxi-caloric coefficient) is approximately equal to 4.821 cal/ml O2 [19]. The estimating of incoming energy by the speed of consumption of oxygen has received the name the method of indirect calorimetry. The energy of molecules of ATP is used basically for the synthesis of molecules, which are included in the organism, also as for other processes requiring expenses of energy (muscle contraction, migration of cells, conduc- tion of nervous impulses, etc.). Regarding the above-described classification of internal variables, the result of the release of energy of molecules of ATP is the emergence of internal variables of the second group, which is accom- panied by excitation of internal variables of the third group. The flux of heat energy from the living system ∆Q is measured by the method of direct calorimetry, that is by direct measurement of the speed of heat production of the organism by means of a calorimeter. The amount of thermal energy coincides with the dissipation of energy that appears at the relaxation of internal variables of the third group with small times of relaxation. The amount of the heat flux is evidence of the presence of quickly relaxating internal variables. In steady state, when there are no changes of thermodynamic functions, incoming and outgoing flows of energy, according to equation (1), are equal. When an organism grows, the incoming stream of energy provokes the emer- gence of internal variables of the second and third group. The relaxation of variables of the third group defines dissipation of energy that in thermal form is deleted completely from the organism. Remind that the internal variables of the first and second groups do not contribute to the dissipation due to their huge relaxation times. Thus, the difference between the incoming and outgoing flows, if exists, gives rise to the emergence of structural variables of the second group, providing the construction of organs and tissues of the organism. Entropy S of the whole organism, due to the mass growth, increases; simultaneously increases in the number of internal variables and their mu- tual connections that is, one can say, increases in structural complexity of 8 the organism. The better, than the total entropy S, characteristic of the living organism is the specific entropy S/M : under the balance of incoming and outgoing energy, specific entropy S/M , as it follows from equation (3), can only decrease. In the situations of growth, the complexity of organism increases, and, so as the greater the complexity, the less entropy, one can assume that the specific entropy does not increase, and it gives (cid:18) S (cid:19) M M S d dt = −σ, σ ≥ 0. (5) At the moment we cannot say that the quantity σ is constant, but later (in Section 3.4), its value will be estimated and appears to be about zero. The permanent exchange of substances and energy between the organ- ism and environment, which is a necessity for the creation of structures and maintenance of functioning, is controlled by the program (the internal instructions) stored with internal variables of the first group. 3.3 Equations of growth From the previous reasoning, it follows that the growth rate of mass of the organism is essentially determined by a difference between quantities of in- coming energy and outgoing heat. Equation (3) shows the linear dependence of the specific growth rate of mass on the difference, which had attracted special attention and attained the name of Ψu-function [5]. The specific psi-function is defined as ψu = 1 M ∆Q ∆t Ξi dξi dt + ηα ∆Nα ∆t . (6) (cid:32) −(cid:88) i (cid:33) (cid:88) α Now one can reformulate relation (3), using definitions (5) and (6), so that the equation of growth can be recorded in the form 1 M dM dt = σ + κ ψu, (7) where κ = M/T S. According to the above equation, the origin of growth is the difference between quantities of incoming and outgoing flows of energy. The growth occurs until the organism will not reach the final steady state defined both by the genetic program, and conditions of the inhabitancy. The majority of poikilothermal animals grow during all life. The homoiothermal animals (mammals and birds) and some of the poikilothermal ones reach the stable mass [20]. 9 In the stationary case the organism should be considered as a system, in which all processes appear balanced: Ξi dξi dt + ηα ∆Nα ∆t = 0. (8) −(cid:88) i ∆Q ∆t (cid:88) α The adult organism, thus, can be considered as the certain constant struc- ture that is being in the non-equilibrium stable state described constant thermodynamic functions. The time dependence of the mass growth for all types of animals can be approximated by a simple equation [21]: dM 1/u dt = V0ct, 0 < c < 1, (9) where V0 is the speed of the growth in initial point in time t = 0; c is a parameter determining the rate of growth; u is a correction factor that is taking into account interfaced processes (morphogenesis, differentiation, etc.). The special case of equation (9) - the equation of Bertalanffy [22] -- can be recorded in the form 1 M dM dt = u ln c ct−t0 − 1 , (10) where t0 is a conditional time, at which assumingly M = 0, c is a factor determining the rate of growth, u is a factor connected with the influence of morphogenetic processes during the growth. 3.4 Testing of the thermodynamic equation of growth To estimate correspondence of equation (7) to reality, we address to results of direct observation of growth of animals in the situations, when one can control the fluxes of energy between the organism and environment [4, 5, 23 - 28]. The fluxes of heat from the organism were measured by the method of direct calorimetry. The fluxes of chemical energy in the organism were estimated by the method of indirect calorimetry according to the rate of oxygen consumption. We can draw certain conclusions from the results. It has appeared, that during an essential period of ontogenesis of organ- isms, the values of incoming and outgoing energy fluxes practically coincide, 10 Figure 1 The dependence of specific growth rate on ψu-function Species: a - yeast Saccharomyces cerevisiae (according to Schaarschmidt et al. [24]; b - larvae of crickets Acheta domestica; c - embryos of lizards Lacerta agilis; d - chicken embryos Gallus domesticus (according to Kleymenov [28]). The depen- dences are approximated by straight lines, according to equation (7). 11 Table of Estimates of Specific Entropy Type (class) κ, g/mW · day σ, day−1 r2 S/M , J/g · K Saccharomyces cerevisiae (Saccharomycetes)1 Acheta domestica (Insecta), larvae2 Lacerta ilis tilia), embryos2 ag- (Rep- Gallus domesticus (Aves), embryos2 (4.73 ± 0.43) · 10−7 (5.70 ± 2.98) · 10−6 0.92 ± 0.08 365 · 106 0.042 ± 0.009 0.011 ± 0.014 0.76 ± 0.16 0.062 ±0.016 0.015 ± 0.046 0.87 ± 0.22 4400 3350 0.166 ± 0.046 0.123 ± 0.098 0.98 ± 0.07 300 Sources of data: 1 -- Schaarschmidt et al. [24]; 2 - Kleymenov [28]. and such a situation has received the name of the current steady state of the organism [4, 7, 18]. However, there is mass growth in such situations, and, according to equation (7), the difference between incoming and outgo- ing flows of energy must exist. One can assume that the growth occurs so slowly, that the difference appears to be so small, that, apparently, cannot be detected with the used methods. Apart of it, when the method of indi- rect calorimetry is used for the assessment of incoming stream of energy, the contribution of processes of glycolysis, which are also processes of synthesis of ATP, are neglected in comparison with the process of oxidizing phospho- rylation. However, the contribution of the processes of glycolysis can appear comparatively essential in this phase of growth. The authentic distinction between data of the indirect (according to the rate of oxygen consumption) and the direct (according to the rate of heat production) calorimetry in situations of the constant environment is observed in cases of the initialization of such processes, as differentiation, morphogenesis etc [28]. At early stages of the growth in ontogenesis, the distinction of results of the direct and indirect calorimetry ψu is great enough to address to an assessment of the adequacy of equation (7). 12 Figure 2 The time dependence of ψu Species: a - yeast Saccharomyces cerevisiae (according to Schaarschmidt et al. [24]); b - larvae of crickets Acheta domestica; c - embryos of lizards Lacerta agilis; d - chicken embryos Gallus domesticus (according to Kleymenov [28]). The depen- dences are approximated by curve lines, according to equation (11). 13 The specific growth rates, as functions of ψu, according to experimental data [24, 28] are shown on Fig. 1. The values of coefficients of correlation and parameters of equation (7) are collected in Table. Though the variability of data is great, the coefficient of correlation between the ψu and specific growth rate differ reliably from zero (p < 0.01). Note that the quantity σ is not significantly different from zero considering the error bars. Thus, specific entropy can be considered as a constant during a significant time interval, and the calculated values of the coefficient κ allow to estimate specific entropy S/M = 1/κ T (see the Table). It is seen that specific entropy decreases in the evolutionary raw: yeast → insects → reptiles → birds. These results show a reduction of entropy or, in other words, an increase of a degree of complexity of living systems during evolution. A positive correlation between the specific growth rate and ψu also has been noted for representatives of different taxons: bacteria Serratta mari- norubta, S. mercescens [27], yeast Sacharamyces cerevisiae [24], larvae of mealworm beetle Tenebrio molitor [25], larvae of house crickets Acheta do- mestica, embryos of sand lizard Lacerta agilis, chicken embryos Gallus do- mesticus [28], juvenile fishes Xiphophorus helleri [23], larvae of axolotl Am- bystoma mexicanum [26], ovogenesis of clawed frogs Xenopus laevis [5]. It is important to have a look at the time dependence of the quantity ψu at the growth of the organisms: it gives an additional way of assessment of adequacy of the equation (7). Comparing equation (10) with the equation (7), we receive: a − b, ψu = 1 − ct−t0 (11) where a = −u ln c/κ; b = σ/κ. On the plots of Fig. 2, the experimental values of psi-function, corresponding to values in Fig. 1, are shown with empty circles. The solid curves are calculated, according to equation (11), at individual values of quantities a, b and c. This comparison confirms the existence of the universal law of growth in the form (10). 4 The Discussion and Conclusions The thermodynamic approach for the open systems, developed by Prigogine [9] and his followers, gives a general foundation for description of dynamics of processes running in living systems. The internal variables, the number of which is huge, describe the deviations of the biological organism, as a thermodynamic system, from equilibrium and appears to be the most essen- tial elements of description. However, we cannot even inventory all internal 14 variables, and are still very distant from the detailed description of the liv- ing organism. We have found that, though the biological organism exists as integral system and all internal variables are anyhow connected with each other, it is useful to allocate some groups of the internal variables according to the relaxation times. The mechanisms of heredity and control, which is the main thing that distinguish the biological organism from a system of coupled chemical reactions, is assumed to be ascribed to a special group (the first group in our classification) of internal variables. The internal variables with relaxation times comparable with the lifespan of organism (the second group) describe the metastable structure that is a real implementation of living organism. The variables of the third group are participating in ther- modynamic processes that are obeying the linear laws of the non-equilibrium thermodynamics. The clarification of functional roles of the internal variables allows us to formulate the thermodynamic equation of growth of organisms that corre- sponds not only to the results of experimental observations but also to basic phenomenological equations of growth. It is remarkable that application of the derived equation of growth to experimental data allows one to estimate the value of specific entropy of living organisms. The specific entropy appears to be the most important thermodynamic characteristic of the living organism. In some examples, it was shown that specific entropy remains constant during the growth and development of the organism. It is possible that this property is valid only for many-cell bodies, when the growth is connected with the multiplication of the high complexity of the original cell. So as the specific entropy does not change substantially during the lifespan, it appears to be an individual characteristic of the or- ganism. The assessments of specific entropy for larvae of crickets, embryos of the sand lizard and chicken embryos confirm that there is decrease of specific entropy and, consequently, an increase of a degree of complexity of the living systems during progressive macroevolution. Appendix: Entropy of the open thermodynamic system The concept of entropy, as a measure of the degraded energy, has arisen at the description of the ideal equilibrium situations. Variation of entropy S under reversible process of a heat transaction to system in the amount ∆Q is defined as ∆S = , (A.1) ∆Q T 15 This circumstance allows to interpret entropy of the systems in equilibrium situations as a measure of thermal, completely degraded energy under some restrictions by a set of constitutive parameters (volume V in the simplest case) that are defined the thermodynamic system. At the extension of con- cept of entropy for non-equilibrium states of the system, it is natural to demand, that interpretation of entropy as a measure of thermal, completely degraded energy is being kept. The ignoring of this requirement has brought to existence of a set of 'non-equilibrium entropies' and appears to be a source of some misinterpretation. We shall demonstrate the realization of such an approach, following the works [13, 14], for general case of open system, which is assumed performs no work. Then, the expression for variation of total internal energy U (the first law of thermodynamics) is recorded in the form of (cid:88) ∆U = ∆Q + (ηα + µα) ∆Nα, (A.2) α where ∆Q is a flux of heat into the system, ∆Nα is a change of substance of a kind α within the system. The enthalpy of unity of the substance of kind α is presented as a sum of two quantities, ηα + µα, one of them µα is chemical potential of the substance. The parity (A.2) is valid both for equilibrium, and for non-equilibrium situations, however, if, in equilibrium states, internal energy U represents completely degraded energy, in a non-equilibrium situation only some part of the total internal energy U , which we shall designate as E, is in the form of thermal energy. Other part of energy is connected with the restrictions caused by constant constitutive parameters (volume V in the simplest case) and relaxating internal parameters ξ. The latter are introduced to describe deviations of the thermodynamic system from the equilibrium state. The total internal energy of the system changes due to the change of the com- position of the system2 and work on variations of the internal parameters ξ, so that it can be recorded as ∆U = ∆E + Ξj ∆ξj + µα ∆Nα. (A.3) j α The part of total internal energy is in the thermal form, and a part in the form that is ready to turn in thermal energy. There are various ways of definition of entropy for the open thermody- namic system in view of all of non-equilibrium processes. In the simplest 2Thanks to W. Muschik, who pointed at the existence of this contribution. 16 (cid:88) (cid:88) version, one keeps definition (A.1), which, apparently, is wrong, because since times of Clausius is known that in non-equilibrium processes for the closed system ∆S ≥ ∆Q T (A.4) In view of it, the definition of non-equilibrium entropy for the closed system should contain additional positive terms ∆S = ∆Q T + .... To define entropy as a measure of thermal energy, we shall return to the law of conservation and transformation of energy for the system, which can be recorded, according to parities (A.2) and (A.3) in the form of ∆E = ∆Q −(cid:88) (cid:88) Ξi ∆ξi + ηα ∆Nα. (A.5) i α Further, adhering to interpretation of entropy as a measure of energy in the form of heat, we follow an equation (A.1) and, neglecting the work of the system, record ∆S = . (A.6) ∆E T Here temperature T is also defined as absolute temperature for the ther- mal part of the system that has thermal energy E. Of course, the real temperature of the system in a non-equilibrium state could differ from this temperature T , but the difference can be considered as an internal variable and included in the list of ξ(inexplicitly). In such a way, equation (A.5) and (A.6) determines the differential of entropy dS = 1 T (cid:32) ∆Q −(cid:88) i (cid:88) α Ξi ∆ξi + (cid:33) ηα ∆Nα . (A.7) Equation (A.7) is the only definition of entropy of the open system, which is consistent with the understanding this quantity as a measure of the degraded energy and empirical evidences. 17 References 1. E.S. Bauer, Theoretical Biology, All-Union Inst. Exp. Med., Moscow-Leningrad, 1935. 2. I. Prigogine, J.M. Wiame, Biologie et thermodynamique des phenomenes irreversibles, Expenentia 2 (1946), 451-453. 3. H.E. Salzer, Toward a Glbbsian approach to the problems of growth and cancer, Acta Biotheor. 12 (1957), 135-166. 4. A.I. Zotin, Thermodynamic aspects of developmental biology, Monographs in Developmental Biology. V. 5, S. Karher, Basel et al., 1972. 159 p. 5. A.I. Zotin, Dissipative structures and Ψu-function, In: Thermodynamics of Biological Processes, Walter de Gruyter, Berlin & NY, 1978, P. 301-304. 6. A.I. Zotin, Changes of ψd and ψu-functions during oogenesis of Xenopus laevis, In Thermodynamics of Biological Processes, Walter de Gruyter, 1978, P. 191-196. 7. A.A. Zotin, A.I. Zotin, Phenomenological theory of ontogenesis, Int. J. Dev. Biol. 41, (1997), 917-921. 8. E. Schrodinger, What is Life? The Physical Aspect of the Living Cell, Cam- bridge Univ. Press, Cambridge, 1944. 9. I. Prigogine, Introduction to Thermodynamics of Irreversible Processes. Sec- ond Revised Edition,. Interscience publ., N.Y.-London, 1961. 10. K.S. Trincher, [Biology and information. Elements of biological thermody- namics], Nauka, Moscow, 1965, 119 p. 11. G. Nicolis and I. Prigogine, Self-Organisation in Non-Equilibrium Systems: From Dissipative Structures to Order through Fluctuations, John Wiley & Sons, New York, 1977. 12. W.B. Cannon, The Wisdom of the Body, W.W. Norton & Company, inc., London, 1932. 13. V.N. Pokrovskii, Extended thermodynamics in a discrete-system approach, Eur. J. Phys. 26, (2005), 769-781. 14. V.N. Pokrovskii, A derivation of the main relations of non-equilibrium ther- modynamics. Hindawi Publishing Corporation: ISRN Thermodynamics 2013, article ID 906136, (2013), 9 p. http://dx.doi.org/10.1155/2013/906136 15. L. Brillouin Science and information theory. Academic Press Inc. Publishers, New York, 1956. 18 16. P.D. Gujrati, Hierarchy of Relaxation Times and Residual Entropy: A Nonequi- librium Approach, Entropy 20 (2018), 149, doi:10.3390/e20030149 17. D. Metzler, Biochemistry: the chemical reactions of living cells, Academic Press, London et al., 1977, 1168 p. 18. A.I. Zotin, R.S. Zotina, [Phenomenological theory of development, growth and aging of organisms], Nauka, Moscow, 1993, 363 p. 19. D.Briedis, R.C. Seagrave, Energy transformation and entropy production in I. Application to embryonic growth, J. Theor. Biol. 110 living systems. (1984), 173-193. 20. A.A. Zotin, [Patterns of growth and energy metabolism in the ontogeny of mollusks], Doctor dissertation thesis. Moscow: IBR RAN, (2009), 334 p. 21. A.A. Zotin, The united equation of animal growth, Amer. J. Life Sci. 3 (2015), 345-351. 22. L. von Bertalanffy, On the von Bertalanffy growth curve, Growth 30 (1965), 123-124. 23. V.A. Grudnickij, [Growth and metabolism in swordtails]. PhD thesis. Moscow: IBR RAN, (1972), 137 p. 24. B. Schaarschmidt, A.I. Zotin, R. Brettel, I. Lamprecht, Experimental inves- tigation of the bound dissipation function: change of the (u-function during growth of yeast, Arch. Microbiol. 105 (1975), 13-16. 25. K.D. Loehr, P. Sayyadi, I. Lamprecht, Thermodynamic aspects of develop- ment for Tenebrio molitor L, Experientia 32 (1976), 1992-1003. 26. V.A. Grudnickij, I.S. Nikol'skaja, [Heat production at early stages of axolotl growth from direct and indirect calorimetry data], Ontogenesis 8 (1977), 80- 82. 27. J. Bermudez, J. Wagensbers, The entropy production in microbiological sta- tionary states, J. Theor. Biol. 1986 (1986) , 347-358. 28. S.Ju. Kleymenov, The mass specific energy metabolism in the early ontogen- esis of animals from the data of indirect and direct calorimetry, PhD thesis. Moscow: IBR RAN, (1996), 147 p. 19
1111.5645
1
1111
2011-11-23T22:49:53
Dynamic Properties of Human Bronchial Airway Tissues
[ "physics.bio-ph", "physics.med-ph" ]
Young's Modulus and dynamic force moduli were measured on human bronchial airway tissues by compression. A simple and low-cost system for measuring the tensile-strengh of soft bio-materials has been built for this study. The force-distance measurements were undertaken on the dissected bronchial airway walls, cartilages and mucosa from the surgery-removed lungs donated by lung cancer patients with COPD. Young's modulus is estimated from the initial slope of unloading force-displacement curve and the dynamic force moduli (storage and loss) are measured at low frequency (from 3 to 45 Hz). All the samples were preserved in the PBS solution at room temperature and the measurements were perfomed within 4 hours after surgery. Young's modulus of the human bronchial airway walls are fond ranged between 0.17 and 1.65 MPa, ranged between 0.25 to 1.96 MPa for cartilages, and between 0.02 to 0.28 MPa for mucosa. The storage modulus are found varying 0.10 MPa with frequency while the loss modulus are found increasing from 0.08 to 0.35 MPa with frequency. The frequency-dependent dynamic force moduli are also compared with different strain rates.
physics.bio-ph
physics
Dynamic Elastic Properties of Human Bronchial Airway Tissues J.-Y. Wanga , P. Mesquidaa,b , P. Pallaic , C. J. Corriganc , T. H. Leec [email protected], [email protected] aDivision of Engineering, King’s College London, Strand, London WC2R 2LS, UK bDepartment of Physics, King’s College London, Strand, London WC2R 2LS, UK cAsthma, Allergy and Lung Biology Division, MRC-Asthma UK Centre in Allergic Mechanisms of Asthma, King’s College London, Guy’s Hospital, London SE1 9RT, UK November 20, 2011 Abstract Young’s Modulus and dynamic force moduli were measured on human bronchial airway tissues by compression. A simple and low-cost system for measuring the tensile-strengh of soft bio-materials has been built for this study. The force-distance measurements were undertaken on the dissected bronchial airway walls, cartilages and mucosa from the surgery-removed lungs donated by lung cancer patients with COPD. Young’s modulus is estimated from the initial slope of unloading force-displacement curve and the dynamic force moduli (storage and loss) are measured at low frequency (from 3 to 45 Hz). All the samples were preserved in the PBS solution at room temperature and the measurements were perfomed within 4 hours after surgery. Young’s modulus of the human bronchial airway walls are fond ranged between 0.17 and 1.65 MPa, ranged between 0.25 to 1.96 MPa for cartilages, and between 0.02 to 0.28MPa for mucosa. The storage mod- ulus are found varying 0.10MPa with frequency while the loss modulus are found increasing from 0.08 to 0.35 MPa with frequency. The frequency- dependent dynamic force moduli are also compared with different strain rates. 1 Introduction Ariway diseases such as asthma are related to the airway wall elasticity. The elastical properties of lung tissue are important to determine the mechanical behoviour of lungs [1]. Over the past few decades, dynamic measurments on the respiratory gas pressure, volume, and flow have been applied on clinical diagonis for understanding the lung mechanics such as spirometry and forced-oscillation technique [2]. The lung resistance Rr and elastance Er are determined by the 1 equation of motion P (t) = Er V (T ) + Rr V (t) + P0 , where P (t) is the time- dependent applied pressure, V is the volume, and V is the flow, and P0 is the pressure corresponding to transpulmonary pressure at end expiration [3]. The dynamical elasticity measurements on airway walls can lead to a better understanding of respiratory ventilation mechanisms and also provide useful information and reference values for clinical tools, such as ultrasound and MRI elastography [Ophir1997,Miller2007]. Stiffness measurements on human and animal trachea rings have been stud- ied in the past two decades [Bert1992,1997,2005]. However, the elastic properties of airway walls from bronchi are little documented. In this study, a simple and low-cost system has been developed for measuring the elastic properties of soft bio-materials. Force-distance measurements on bronchial wall segments from lung cancer patients with COPD have been performed. 2 Basic Principles Determining Young’s Modulus from Force-Distance Curve The compression method of loading and unloading is based on the fact that the displacements recovered during unloading are largely elastic, in which case the Young’s modulus, E , can be simply analyzed from the unloading curve where the Hooke’s law holds [4]. Young’s modulus of an elastic material is a ratio of uniaxial stress (σ) over strain (ǫ) which can be related to an external force F over the uniaxial length changing ∆L by E = ǫ σ = F /A0 ∆L/L0 = F ∆L · L0 A0 (1) where L0 is the initial length and A0 is the initial cross-sectional area of the material without any applied force. In this study, taking the initial unloading process into estimation, the Young’s modulus can be given by E = dF dh · L0 − ∆L A0 + ∆A = S · L0 − hf (w + ν hf )2 (2) where (L0−hf ) is the material length at initial unloading process while (w+µhf ) is the material width at initial unloading process with Poisson’s ratio ν (ranged between 0.47 to 0.5 for biomaterials [5, 6, 7]). The elastic unloading stiffness, S = dF /dh, is defined as the slope of the unloading curve during the initial stages of unloading. A schematic illustration of force-displacement curve is shown in figure 1 where one cycle of loading and unloading of an indenter on a material (which is relatively soft to the indenter) is presented [4]. Dynamic Force Moduli 2 F e c r o F loading S = dF/dh unloading Displacement h h f Figure 1: Schematic illustration of a cycle of loading and unloading force- displacement curve. The unloading stiffness S is defined as the slope of the force and displacement at the initial unloading process. hf is the final displace- ment of the material [4]. The stress-strain relationship for a linear viscoelastic material under a sinu- soidal loading can be expressed as [8, 9] σ0 cosω t = ǫ0E ′ cosω t + E ′′ sin(ω t + φ) (3) where σ0 is the stress amplitude, ǫ0 is the strain amplitude, ω is the angular frequency, φ is the phase between the stress and strain. The storage modulus, E ′ , which corresponds to the in-phase term, can be given by [Herbert2008, RA Jones] E ′ = cosω t σ0 ǫ0 (4) (5) (6) (7) and the loss modulus, E ′′ , the out-of-phase term, is given by E ′′ = σ0 ǫ0 sinω t. The stress and strain can also be expressed in a complex form, i.e., σ∗ = σ0 eiωt ǫ∗ = ǫ0 eiωt+φ . The complex form of the modulus E ∗ can thus related to the E ′ and E ′′ by E ∗ = σ ǫ = σ0 ǫ0 eiφ = σ ǫ (cosφ + sinφ) = E ′ + iE ′′ . 3 System The system developed for measuring the elastic properties of small biomaterials is shown in figure 2. A loud speaker (4, BG 10, 8 Ω , Visaton, Germany) is used as an actuator for driving the compressing rod. The compressing rod is made of 3 Loud speaker T D V L g n i s s e r p m o c d o r Sample Force Sensor Figure 2: The system diagram for measuring the elastic modulus. The com- pressing rod with flat end is driven by a dynamic loud speaker. The distance is measured by a LVDT (linear variable distance transducer). The sample is placed on top of the force sensor which monitors the force applied on the sample. stainless steel with a flat end of diameter 1.25 cm , and fixed on the diaphram of the speaker. The loud speaker is powered by a modified stereo amplifier (K4003, Velleman-Kit, Beigium) and the input signal is controlled by a computer with LabView programmes and an interfacing DAQ card . A liner variable distance transducer (LVDT, Position sensor M-0.5, Applied Measurements Ltd., UK) is used for measuring the displacement of the compressing rod which covers a linear distance of ± 0.5 mm. The digital noise from the LVDT is approximately 4.05 mV which corresponds to 5.05 µm. A force sensor modified from a balance (full-scale range 50g, resolution 0.01g, APS-50, Farnell, UK) is placed under the compressing rod. The digital noise from the force sensor is approximately 4.02 mV which corresponds to 0.09 mN. The sample is placed on the force sensor and under the compressing rod. Two different types of silicone rubber (2-mm thick, 4 mm × 4 mm cross- sectional area) have been tested with the system which have Young’s Modulus of 0.32 MPa (794N, DowCorning, USA) and 1.80 MPa (Bond Flex 100HMA, Bostik, UK). The system gives values within 3 % of the published values from the silicone sample datasheets at 21 oC. 4 Experiment Setup The human bronchial airway tissues were obtained from the surgery-removed lung specimens donated by patients with lung cancer and COPD. All the exper- iments in this study were under the human tissue act policy in King’s College London and approved by the ethical committee in Guy’s and St Thomas’ NHS 4 (a) Airway wall (b) Cartilage R1 R2 R3 2 1.5 1 0.5 ) a P M ( E 2 1.5 1 0.5 R2 R3 (c) Mucosa R1 R2 R3 2 1.5 1 0.5 0 0 0.05 0.1 0.15 strain 0.2 0.25 0.3 0 0 0.1 strain 0.2 0.3 0 0 0.1 strain 0.2 0.3 Figure 3: The Young’s Modulus of (a) bronchial airway wall, (b) cartilage and (c) mucosa from three cancer patients R1-3 are plotted against strain. hospital. Samples including bronchial airway walls, cartilages, and mucosa were dissected from the tissues and preserved in PBS solution at room temperature (21 ± 1 oC). Measurements were performed within 4 hours after the surgery at room temperature. The dissected specimens were cut into roughly square shape with a cross sectional area of about 2 mm × 2 mm and placed under the com- pressing rod with flat end (1.25 cm diameter) and the sample surface can be fully in contact with the rod during the measurements. Young’s modulus is estimated from the force-distance curves which the mea- surements were taken under a constant speed of 0.1 Hz. The siffness of the instrument was calibrated by testing without any sample before each measure- ment and the displacement from the instrument was subtracted from the force- distance curve to give the net displacement of the sample. The dynamic force measurements were taken under low frequencies (3-45 Hz). The compressing rod was loaded on the sample with knowing force and displacement before performing the frequency sweep to make sure the sample was fully incontact with the rod. The input sinusoidal signal with 5-mV ampli- tude was powered by the computer corresponging to 4.51 mN force. The phase between the force and displacement signals was recorded for each frequency with a LabView programme under 1 kHz sampling rate. Calibrations on frequency dependent stiffness of the instrument has also been done with frequency sweep at certain load and subtracted from the dynamic force measurements. 5 Results Young’s modulus with different strain rates of the bronchial airway tissues from three cancer patients are shown in figure 3. Results show the Young’s modulus of the airway wall ranges between 0.1 and 1.0 MPa (with strain from 0.05 to 0.22), and of cartilage ranges between 0.2 and 2.2 MPa (with strain from 0.05 5 (a) R3 bronchial airway, d=0.85±0.15 (cm) 750 e g a r o t s − ) a P k ( ’ E 700 650 600 550 500 450 400 350 300 250 0 airway wall, strain 5.41 % cartilage, strain 5.66 % mucosa, strain 4.38 % 10 20 30 40 50 f (Hz) (b) R3 bronchial airway, d=0.85±0.15 (cm) s s o l − ) a P k ( " E 500 450 400 350 300 250 200 150 100 50 0 0 airway wall, strain 5.41 % cartilage, strain 5.66 % mucosa, strain 4.38 % 10 20 30 40 50 f (Hz) Figure 4: The dynamic force modulus E’ and E” of bronchiole tissue (with diameter 0.85 ± 0.15 cm) from one cancer patient. The frequency-dependent modulus were compared between the bronchial airway wall, the cartilage, and the mucosa at similar preloaded strains of 5 ± 0.6 %. (a) R2 & 3 Cartilage, E’ 1300 R2 strain 10.33% R2 strain 7.75 % R3 strain 12.02 % R3 strain 5.66 % e g a r o t s − ) a P k ( ’ E 1200 1100 1000 900 800 700 600 500 400 300 (b) R2 & 3 Cartilage, E" R2 strain 10.33 % R2 strain 7.75 % R3 strain 12.02 % R3 strain 5.66 % 600 550 500 450 400 350 300 250 200 150 100 s s o l − ) a P k ( " E 200 0 5 10 15 20 30 35 40 45 50 25 f (Hz) 50 0 5 10 15 20 30 35 40 45 50 25 f (Hz) Figure 5: The dynamic force modulus E’ and E” of bronchial cartilage samples from two cancer patients (R2 and R3). The frequency-dependent modulus were compared at different preloaded strains with the same oscillation amplitude. 6 to 0.15), whereas mucosa of the smallest values ranged betwen 0.35 and 0.01 MPa (with strain between 0.05 to 0.26). The dynamic force mudulus E’ and E” of bronchial airway wall, cartilage, and mucosa from one patient is plotted in figure 4 with similar preloaded strain. The storage modulus of the airway wall and mucosa samples is found increasing with frequency while decreasing in the cartilage sample with the strain of 5±0.6%. Figure 5 shows the dynamic moduli E’ and E” of bronchial cartilage samples from two cancer patients with different preloaded strains. The storage modulus is found increasing slightly with the frequency in one patient while decreasing with frequency between 10-45 Hz in the other patient. Acknowledgement 1 We thank Ms Juliet King, Dr Paul Cane, Dr Emma McClean and Dr Amanda Murphey, from Guy’s and St Thomas’ Hospital for patient recruitment and tissue dissecting. We also thank the Biomedical Re- search Centre in Guy’s and St Thomas’ NHS Trust, London, UK for funding this project. References [1] Faffe DS, Zin WA. Lung Parenchymal Mechanics in Health and Disease. Physiol Rev 89 : 759-775, 2009. [2] Zera F, Lorino A, Lurino H, Harf A, Macquin-Mavier I. Forced Oscillation Technique vs Spirometry to Assess Bronchodilatation in Patients With Asthma and COPD. Chest 108 : 41-47, 1995. [3] Navajas D, Maksym GN, Bates JHT. Dynamic viscoelastic nonlinearity of lung parenchymal tissue. J Appl Physiol 79 : 348-356, 1995. [4] Oliver WC, Pharr GM.On the generality of the relationship among con- tact stiffness, contact area, and elastic modulus during indentation.J Mater Res 7 : 613-617, 1992. [5] Fung YC. Biomechanics: Mechanical Properties of Living Tissues. New York: Springer-Verlag, 1981. [6] Chen EJ, Novakofski J, Jenkins WK, O’Brien WD Jr. Young’s mod- ulus measurements of soft tissues with application to elasticity imaging. IEEE Trans Ultrasonics Ferroelectrics, and Freq Contrl 43 : 191-194, 1996. [7] Toyras J, Lyyra -Laitinen T, Niinimki M, Lindgren R, Nieminen MT, Kiviranta I, Jurvelin JS. Estimation of the Young’s modulus of ar- ticular cartilage using an arthroscopic indentation instrument and ultrasonic measurement of tissue thickness. J Biomech 34 : 251-256, 2001. [8] Herbert EG, Oliver WC, Pharr GM. Nanoindentation and the dy- namic characterization of viscoelastic solids. J Phys D: Appl Phys 41 : 1-9, 2008. 7 [9] Herbert EG, Oliver WC, Pharr GM. Measurement of hardness and elastic modulus by instrumented indentation: Advances in understanding and refinements to methodology. J Mater Res 19 : 3-20, 2004. 8
1501.06742
1
1501
2015-01-27T11:04:34
Free energy landscape and characteristic forces for the initiation of DNA unzipping
[ "physics.bio-ph", "cond-mat.soft", "q-bio.BM" ]
DNA unzipping, the separation of its double helix into single strands, is crucial in modulating a host of genetic processes. Although the large-scale separation of double-stranded DNA has been studied with a variety of theoretical and experimental techniques, the minute details of the very first steps of unzipping are still unclear. Here, we use atomistic molecular dynamics (MD) simulations, coarse-grained simulations and a statistical-mechanical model to study the initiation of DNA unzipping by an external force. The calculation of the potential of mean force profiles for the initial separation of the first few terminal base pairs in a DNA oligomer reveal that forces ranging between 130 and 230 pN are needed to disrupt the first base pair, values of an order of magnitude larger than those needed to disrupt base pairs in partially unzipped DNA. The force peak has an "echo," of approximately 50 pN, at the distance that unzips the second base pair. We show that the high peak needed to initiate unzipping derives from a free energy basin that is distinct from the basins of subsequent base pairs because of entropic contributions and we highlight the microscopic origin of the peak. Our results suggest a new window of exploration for single molecule experiments.
physics.bio-ph
physics
Free Energy Landscape and Characteristic Forces for the Initiation of DNA Unzipping Ahmet Mentes*,(cid:91), Ana Maria Florescu¶ (cid:91), Elizbeth Brunk†, Jeff Wereszczynski‡, Marc Joyeux(cid:93), and Ioan Andricioaei*, § *Department of Chemistry, University of California, Irvine, CA 92697 ¶Max-Planck Institute for the Physics of Complex Systems, Dresden, Germany, and Interdisciplinary Research Institute, Universit´e des Sciences et des Technologies de Lille (USTL), CNRS USR 3078, 50, Avenue Halley, 59568 Villeneuve d'Ascq, France †Fuels Synthesis Division, Joint BioEnergy Institute, Emeryville, CA 94608, and Department of Chemical and Biomolecular Engineering, Department of Bioengineering, ‡Department of Physics, Illinois Institute of Technology, 3300 South Federal Street, University of California, Berkeley, CA 94720 (cid:93)Laboratoire Interdisciplinaire de Physique, Universit´e Joseph Fourier- Grenoble 1, BP 87, Chicago, IL 60616 38402 St Martin d'Heres, France §Email: [email protected] (cid:91)equal contribution October 10, 2018 Abstract DNA unzipping, the separation of its double helix into single strands, is crucial in modulating a host of genetic processes. Although the large-scale separation of double-stranded DNA has been studied with a variety of theoretical and experimental techniques, the minute details of the very first steps of unzipping are still unclear. Here, we use atomistic molecular dynamics (MD) simulations, coarse- grained simulations and a statistical-mechanical model to study the initiation of DNA unzipping by an external force. The calculation of the potential of mean force profiles for the initial separation of the first few terminal base pairs in a DNA oligomer reveal that forces ranging between 130 and 230 pN are needed to disrupt the first base pair, values of an order of magnitude larger than those needed to disrupt base pairs in partially unzipped DNA. The force peak has an "echo," of approximately 50 pN, at the distance that unzips the second base pair. We show that the high peak needed to initiate unzipping derives from a free energy basin that is distinct from the basins of subsequent base pairs because of entropic contributions and we highlight the microscopic origin of the peak. Our results suggest a new window of exploration for single molecule experiments. Key words: DNA denaturation; molecular dynamics; coarse-grained force fields; umbrella sam- pling; single-molecule force spectroscopy; Brownian dynamics Initiation of DNA mechanical unzipping 2 1 Introduction Many essential genetic processes, such as replication, transcription, recombination and DNA repair, involve unzipping of double-stranded DNA (dsDNA) by proteins that disrupt the hydrogen bonds between complementary bases on opposite strands (1). The detailed understanding of the nature of DNA mechanical separation dynamics, and of the energetics and forces for the conformations that occur during unzipping is also relevant for single-molecule DNA sequencing; high-resolution measurements of the forces may lead to novel ways of sequencing DNA by providing the ability to read the base identities from the distinct signatures resulting when separating the different types of base pairs (2). Moreover, single molecule studies in which DNA is being pulled on via atomic force microscopy (AFM) (3), by optical (4) or magnetical (5) tweezers, or is unzipped through nanometer-sized pores (6) are particularly useful to gauge the mechanical response to external stimuli. Such insights into DNA elasticity (7) and the resilience of its double strand to unzipping can provide useful information for the design of nanomechanical devices constructed of DNA (8) and for the build-up of molecularly engineered DNA scaffolds for molecular-size electronics or for crystalline-state biomolecules that otherwise would be impossible to crystalize (9). Recent experiments performed by pulling dsDNA apart with a constant force (10 -- 14) show that dsDNA separates into ssDNA (single-stranded DNA) when the applied force exceeds a critical value Fc ∼ 12 pN. Moreover, for forces near Fc, the dynamics of the unzipping process is highly irregular. Rather than a smooth time evolution, the position of the unzipping fork progresses through a series of long pauses separated by rapid bursts of unzipping (11). However, because of their low spatial resolution, single molecule techniques cannot yet reveal the first steps of opening a fully base-paired double helix from a blunt end, e.g., the opening of the terminal base pair. For AFM, for example, typical force constants of the cantilever are in the 10 − 20 pN/A range, which, using equipartition arguments, yields fluctuations on the scale of several A and the best resolution that can be reached via AFM is currently estimated to be on the order of 10 base pairs (15 -- 17). As a consequence, unzipping straightforwardly only the first few base pairs of the sequence can not be achieved in current pulling experiments. The opening of terminal base-pairs in blunt-end duplexes is important in initiating DNA melting (18, 19). It is also a biologically important step in the action of nucleic acid processing enzymes (20) and in nucleic acid end recognition by retroviral integrases (21). Moreover, the DNA replication process is a good example of instances where double strand DNA must be unzipped mechanically by polymerases (22). Concomitantly with the understanding that terminal base pair opening is biologically relevant, it is important to note the fact, well-established experimentally and compu- tationally, that the first base pair frays naturally, and that it exists in a relatively fast equilibrium between paired and unpaired or frayed states. While this equilibrium is fast compared to the time scale of the pulling apparatus that could be used in single molecule experiments to probe the un- zipping, it is, at the same time, slow relative to the capabilities of all-atom molecular dynamics simulations. The fraying of first base-pairs has been studied in NMR experiments (23, 24); this provided estimates for the equilibrium and kinetic constants for the paired-frayed conformational transition. Themodynamical data was consistent with the view that frayed states are unfavor- able enthalpically due to loss of stacking stabilization, but that they are stable entropically. The experimental estimates for the populations of the frayed state were in the 10-30% range for CG pairs, and up to 50% for AT pairs, with ample variance depending on experimental conditions. A recent simulation study (19) has additionally provided the atomistic details of terminal base-pair fraying. The kinetics of fraying has been also investigated, with experimental reports concluding that this process is faster than 1 ms (25), while in a computational study of multiple free base pair spontaneous stacking/unstacking in aqueous solution at 310 K transitions occured on a time-scale Initiation of DNA mechanical unzipping 3 of 10 ns (26). In the theoretical arena, the fundamentals of DNA denaturation has been studied since the 1960s and several generations of models for DNA unzipping have been developed (27 -- 31). Arguably the most popular ones are the Peyrard-Bishop (PB) model (32) and its extension, the Peyrard-Bishop- Dauxois (PBD) model (33) with an extra term in stacking to better reproduces experimental data. They have been extensively used to describe DNA thermal denaturation (34) or the dynamics of pulling DNA by an external force (35). Other models have been developed to investigate quantita- tively the difference between DNA unzipping by force and thermal or fluctuation-induced melting of dsDNA (28), or for studying interactions between two single DNA strands (30). These modeling approaches, while revealing the fundamental statistical mechanical picture, are not detailed enough to capture all intricacies of unzipping. An important advancement has been recently made via a semi-microscopic theory of DNA mechanical unzipping advanced by Cocco et. al. (36, 37). This theory accounts for hydrogen bonds, stacking interactions and elastic forces to investigate experimentally observable aspects of DNA unzipping by externally applied forces or torques. Quite interestingly, the calculation of the forces needed to keep the two extremities of the dsDNA molecule separated by a given distance lead in this model to the prediction for the existence of an extremely large force barrier that opposes initial double helix unzipping, namely a ∼ 250 pN force peak occuring at ∼ 2 A separation from the equilibrium base-pair distance (37). This is remarkable because it is more than an order of magnitude larger than the unzipping forces for DNA in the bulk (i.e., forces averaged over scores of unzipped basepairs) previously measured in the various experimental settings. Because both analytically solvable models and experiments can only reveal a limited number force and extension), it is crucial that they are complemented by all-atom of observables (e.g. simulations. This allows one to better understand the dynamics and observe the microscopic effects that pulling forces have on all degrees of freedom and physical properties of the system (38). A previous atomistic molecular dynamics study of DNA mechanical denaturation (39) focused on the sequence effects that occur during non-equilibrium DNA unzipping (with pulling speeds orders of magnitude larger than those in single molecule tweezer experiments) and not on the equilibrium forces needed for the initiation step. The authors observed jumps and pauses in denaturation which they attribute to the inhomogeneity of the DNA sequence they have used: AT (adenine-thymine) rich regions melt earlier (that is at smaller forces) than GC (guanine-cytosine) rich regions because AT base pairs contain two hydrogen bonds whereas GC base pairs contain three hydrogen bonds. Initiation of DNA mechanical unzipping 4 Figure 1: Schematic representation of the DNA sequence and the forces applied for unzipping: bases G, C, A and T are shown in yellow, orange, blue and magenta, respectively and the bonds forming the base-pairs are shown with a dotted line. Forces were applied on two backbone atoms of the first base pair: the O3(cid:48) atom of the C1 residue at the 5' end of one strand and on the O3(cid:48) atom of the G12 residue at the 3' end of the other strand in the all atom simulations, and on the pseudo-atoms describing the sugars of the same residues in the coarse-grained simulations. The terminal C1-G12 base pair is an example of what we refer to in the text as the "first" base pair that needs a higher force to be unzipped in comparison to the subsequent base pairs. The purpose of the present study is to provide a better understanding of the onset of DNA denaturation and an exploration of the origin of any unusually high forces that occur at the very early stage of unzipping (that is the opening of the first 1-2 base pairs) via detailed molecular simulations and subsequent theoretical analysis. To this end, the rest of the paper is organized as follows. First, we compute, along the base separation coordinate, mean force and free energy profiles of a dodecamer of helical B-DNA with the base sequence d(CGCAAATTTCGC)2 using molecular dynamics simulations, umbrella sampling and the weighted histogram analysis method (WHAM) to obtain and atomically detailed potential of mean force profile. Then, in addition to atomistic calculations, to explore the presence of force peaks in simulations with lower (mesoscopic) levels of details and to extend the range of DNA sequences studied, we also perform simulations using a coarse-grained DNA model with three sites per nucleotide (40). Lastly, we also show that we can derive this force peak analytically in the formalism of the Peyrard-Bishop model (32). Taken collectively, our simulation and analysis results reveal that the opening of the first DNA base pair needs significantly larger forces than the opening of the subsequent ones not only to break the hydrogen bonds that form that base pair, but also to overcome an entropic barrier due to stacking interactions. Additionally we reveal that a'second order' contribution to the force peak stemming from non-native H-bonds (i.e., between base pairs that were not originally H-bonded in the intact dsDNA) exists, and that, concomitantly to the development of the force peak, a peak in the torque about the DNA axis develops upon initial unzipping. AGCTH-bondsFORCEFF5`3`FF3`5`FF5`3` Initiation of DNA mechanical unzipping 5 2 Simulation Methods 2.1 All-Atom Molecular Dynamics Simulations We simulated at atomic detail the first steps in the mechanical denaturation of a dodecamer of helical B-DNA with the sequence d(CGCAAATTTCGC)2. Fig. 1 depicts schematically the DNA sequence and the forces that are exerted to induce mechanical unzipping.; it displays the four nucleotides G, C, A and T and the hydrogen bonds between the base pairs. The same labeling and coloring strategy as in Fig. 1 is followed for the other figures of this article. The external forces applied to initiate unzipping are also shown. In the atomistic simulations they were applied on the two O3(cid:48) atoms of the first base pair, i.e., the O3(cid:48) atom of the C1 residue at the 5' end (i.e., the first Cyt residue on one strand) and the O3(cid:48) atom of the G12 residue at the 3' end of the other strand. (Numbering is such that bases are labeled 1 through 12 from the 5' to the 3' direction on both strands, see also Fig. 6). To account for the natural fraying of terminal base pairs, two starting geometries were used in the simulation: a paired first base pair and a frayed one. For the frayed first base pair case, the initial equilibrium distance between O3(cid:48) atoms pulled apart is the same as that for the paired case, and the C1 residue at 5' end (on DNAs1-DNA strand1) and the G12 residue at 3' end (on DNAs2- DNA strand2) were flipped out of the backbone by alterations of the corresponding dihedral angles. The structures such prepared then underwent molecular dynamics simulations using version 34 of the CHARMM software package (41), with the CHARMM27 nucleic acid parameters (42, 43). The reaction coordinate was defined as the separation between the C and G O3(cid:48) atoms of residues 1 and 12, respectively, and we applied harmonic constraints to the separation distance ρ of these atoms. The functional form of the potential used was ku(ρ−ρ0)2. Using umbrella sampling trajectories, statistical data for free energy calculations was collected during the structural changes along the coordinate. The selected atoms were harmonically restrained such as to maintain a separation within approximately 1 A of the specified equilibrium distance ρ0, ranging from 14.50 A, which is the base-pairing equilibrium value, to 30.00 A (which corresponds to 15.5 A separation from base-pairing equilibrium) with increments of 0.05 A for the first 4.5 A separation and 0.25 A for the rest of the windows. The last configuration of the trajectory in each window was used as the initial condition for the next window. A total of 130 windows were calculated, each running for 800 ps. Each 800 ps trajectory was then post- processed using WHAM (44, 45). The more numerous umbrella sampling windows for the first 4.5 A we generated for better resolution in the initial free energy profile. We used the explicit solvent TIP3P potential for water (46). The DNA structure was overlaid with a water box previously equilibrated at 300 K, with dimensions 56 A × 56 A × 56 A, and was initially aligned so that the DNA molecule's primary axis is parallel to the x-axis. The water box contained 3362 TIP3P water molecules, and 22 sodium ions, needed to make the solution electrically neutral. Periodic boundary conditions were used and electrostatic interactions were accounted for using the Particle-Mesh Ewald Method (47), with a real-space cutoff at 12.0 A for non-bonded interactions. The leapfrog Verlet algorithm was used with Nos´e-Hoover dynamics (48, 49) with a coupling constant (thermal inertia parameter) of 50 internal (AKMA) units (50) to keep the temperature constant at 300 K throughout the simulations. The system underwent 100 steps of steepest-descent minimization followed by 1000 steps of the adaptive-basis Newton-Raphson minimization. It was then heated to 300 K over an equilibration period of 800 ps with harmonic restraints applied to the O3(cid:48) atoms in order to prevent the helical axis from becoming unaligned with the z-axis, restraints which are then gradually removed during the production runs. The SHAKE algorithm (51) was used to constrain all covalently-bound hydrogen atoms. The biased umbrella sampling trajectories were post-processed with the weighted histogram Initiation of DNA mechanical unzipping 6 analysis method (WHAM) (44) (as implemented in Ref. (45)), in order to obtain the unbiased free energy values as well as thermodynamic quantities from an unbiased system. Error bars were calculated with Monte Carlo bootstrap error analysis, with repeated computations of the average of resampled data and calculation of the standard deviation of the average of the resampled data. The later is an estimate for the statistical uncertainty of the average computed using the real data. Because for separation distances below 4.5 A the energy is averaged over more windows than for higher separations, this leads to the smaller error bars in that region of interest. The force along the reaction coordinate was computed as the derivative of W (ρ) with respect to ρ, F (ρ) = −dW/dρ and is rigorously (52) the canonical-ensemble thermodynamical average of the force needed to keep the two strands separated by a distance ρ. To determine the vectorial force components, the mean force along the Cartesian x, y and z axes was computed by taking the derivative of W (ρ) with respect to xi, yi and zi of the i-th atoms involved, , (1) (cid:104)fxi(cid:105) ≡ −(cid:104) dW dxi (cid:105) = − dW dρ (cid:104) dρ dxi (cid:105) = − dW dρ ((cid:104)xi(cid:105) − xref ) ρ · N with the corresponding permuted expressions for the y and z directions. We were also interested to compute any torque τ to which DNA is subjected upon increasing the separation distance by finding the backbone vector forces (on atoms C1:O3(cid:48) and G12:O3(cid:48)) in the x, y and z directions, and performing the cross product with the radii vectors of the DNA helix (53). For example, the torque in the helical z-direction is (cid:104)τzi(cid:105) = − dW dρ · (xı + y) × (((cid:104)xi(cid:105) − xref )ı + ((cid:104)yi(cid:105) − yref )) ρ · N , (2) with similar expressions for the x and y directions. 2.2 Coarse-grained simulations For coarse-grained simulations we used the three-site-per-nucleotide DNA model developed by Knotts et al. (54) with the parametrization described in Ref. (40). In this model, each nucleotide is mapped onto three interaction sites (beads): one for the phosphate, one for the sugar ring and one for the base. The equilibrium positions of the three beads are derived from the coordinates of the atoms they replace as follows: for phosphates and sugars, the bead is placed in the center of mass of the atomic structure of the respective group, for adenine and guanine it is placed on the position of the N1 atom of the geiven base, while for cytosine and thymine it has the coordinates of the N3 atom of the given base. The interaction potential between these beads comprises six terms: Epot = Vbond + Vangle + Vdihedral + Vstack + Vbp + Vqq, (3) with Vbond, Vangle and Vdihedral the bonded contributions (stretch, angle bending and torsion re- spectively), and base stacking (Vstack), base pairing (Vbp) and electrostatic interactions (Vqq) -- the Initiation of DNA mechanical unzipping 7 non-bonded terms -- described by: Vbond = k1 Vangle = Vdihedral = kφ Vstack =  (cid:88) (cid:88) i i<j (cid:88) (cid:88) (cid:88) kθ 2 i − 2 i ij rij (di − d0 i )4 (di − d0 i )2 + k2 i (θ − θ0 i )2 [1 − cos(φi − φ0 i )] (cid:88) (cid:88)  (cid:32) r0 (cid:33)12 (cid:33)6 (cid:32) r0 5 (cid:32) r0 (cid:32) r0 (cid:33)12 (cid:33)10 5 (cid:32) r0 (cid:33)12 (cid:33)10 (cid:32) r0 (cid:17) exp −(cid:16) rij (cid:88) − 6 − 6 ij rij ij rij ij rij ij rij + 1 ij rij κD , 4πH2O i<j rij Vbp = AT +GC AT base pairs GC base pairs e2 Vqq =  +  + 1 + 1 where di denotes the distance between two beads connected by the bond i, θi is the angle between three consecutive sites on the same strand and φi is the dihedral angle defined by four consecutive beads (also along the same strand). In the non-bonded terms, ri,j is the distance between sites i and j. In all equations the values with the superscript index 0 are equilibrium values for the respective quantities. For their numerical values we refer the reader to reference (54). The force constants for the bonded terms are the following: k1 = 0.26 kcal·mol−1A−2, k2 = 26 kcal·mol−1A−4, kθ = 104 kcal·mol−1, kφ = 1.04 kcal·mol−1. The stacking interactions act between the first and second nearest neighbors and  = 0.26 kcal·mol−1. The base pairing term acts only between native pairs, with AT = 3.90 kcal·mol −1 and GC = 4.37 kcal·mol−1. Finally, electrostatic interactions are considered to occur only between phosphates, which carry one elementary charge each. In the expression of the Debye-Huckel potential e is the electron charge, H2O = 780 is the dielectric constant for water at room temperature expressed as a function of the dielectric permittivity of vacuum, and rD = 13.603 A is the Debye length for 50 mM Na+ ion concentration. Compared to Ref. (54), we use different values AT and GC, and exclude non-native base pairing; this was done because the original set of parameters can sometimes cause the formation of two pairing bonds per base, a fact which induced melting temperatures that were too high. Consequently we modified the pairing energy values to correctly describe thermal and mechanical denaturation. We simulated the mechanical unzipping of the dodecamer simulated in the atomistic model, d(CGCAAATTTCGC)2, plus two other dodecamers, d(TGCAAATTTCGC)2 and d(CTCAAATTTCGC)2 in which we changed the first, and respectively, the second base pair from CG to AT. Dynamics was propagated by integrating Langevin's equations: mj d2rj dt2 = −∇Epot − mjγ drj dt +(cid:112)2mjγkBT ξj(t) where mj is the mass of site j, rj is its position vector, with the friction coefficient γ and the Gaussian white noise ξj(t) obeying fluctuation-dissipation: (cid:104)ξi(t)ξj(t(cid:48))(cid:105) = δi,jδ(t − t(cid:48)) (4) (5) (6) Initiation of DNA mechanical unzipping 8 The first term on the right hand side of eq. 5 denotes the forces resulting from the potential, the second one describes the friction due to the solvent, while the third one is a thermal random noise. We integrated the equations of motion using a second order algorithm with a time step of 10 fs and a friction coefficient γ of 5 ns−1. A detailed discussion of the choice of γ and of its influence on the simulation results can be found in reference (40). The temperature was set to 293 K. We have modeled mechanical unzipping of the DNA sequence by pulling apart with a constant rate the sugar groups that are part of the first base pair. We computed then the average of the projection along the separation axis of the internal forces acting on the two beads. For each point the force was averaged over 107 time steps (0.1 µs), corresponding to an increase in separation distance of 0.1 A. We have previously used this approach and validated the model and its parameters with respect to DNA unzipping; we refer the reader to Ref. (40) for the details of this validation. 3 Results Using the computational techniques presented in the Simulation Method section, we have performed equilibrium studies of the forces required for the mechanical unzipping of short DNA sequences. We have simulated pulling by the first base pair up to a separation distance of 14 A from equilibrium using all-atom molecular dynamics (CHARMM), and up to 50 A using the coarse-grained model described above. The analysis of the simulation results presented here is mainly focused on what happens at small separation distances, that is, corresponding to the opening of the first two base pairs. To our knowledge, this is the first time that the onset of DNA mechanical denaturation is studied in such detail. From the all-atom simulations we computed a free energy profile (the potential of mean force (PMF)), whose derivative with respect to the separation coordinate was used, according to the definition of the PMF (52)) to obtain the average force needed to keep the first base pair open at any given separation. Because terminal DNA base pairs are known to fray, we had to use two sets of initial conditions, one base-paired, one frayed (as described in the Introduction section). Additionally, since the fraying/unfraying equilibrium is established on time scale much shorter than that of the pulling aparatus used in single molecule pulling (typically A per ms), during a typical pulling one expects to experience time averaging between the two states, so pulling would give the weighted mean (e.g., 30-70% or 10-90%) of the fully paired and fully frayed profiles. The free energy profile along the separation coordinate is shown in Fig. 2. Over a baseline of increasing free energy as a function of separation (whose constant slope averages to the value of the minimum bulk force needed to unzip DNA), we observed significant "pits" or free energy. They introduce higher slopes in the profile, and, since the slopes are proportional to the magnitude of the mean force, they are responsible for the larger forces for the separation of the first and second base pairs (see Figure 3). We also computed conformational entropies (55) at each separation distance using the quasi-harmonic analysis method (56), in which quasi- harmonic frequencies were calculated from diagonalizing the mass-weigthed covariance matrix of nucleic acid atomic conformational fluctuations in each umbrella sampling window. The calculated conformational entropy profile around the first base pair unzipping is shown in the inset to Fig. 2. We observed a substantial conformational entropy contribution to the "ant-lion pit in the free energy profile along the first base-pair unzipping distance, pointing at a large entropic contribution to the initial slope of the free energy and hence to the force peak. Initiation of DNA mechanical unzipping 9 Figure 2: Free energy of DNA unzipping from all-atom simulations with various frayed−paired populations for the first base pair. First base-pair fully paired (black), fully frayed (green), 10% − 90% frayed−paired (blue), 30%− 70% frayed−paired (pink). Arrows point to onset of first and the second base pair unzipping. The significantly deeper, "antlion-pit" free energy well for first base pair separation (between 0-2 A) is a key feature explaining a steep increase in the force required for initial unzipping (see main text). Inset: The conformational entropy contribution to the the free energy profile along the separation distance for unzipping the first base-pair. From the coarse-grained simulations we computed directly the corresponding force at each distance by equilibrium averaging of the force needed to keep the distance between the phosphates of the first base pairs at a given separation. We focus on the analysis of the force peak obtained via these simulations in this section. Before that, we start with an analytical computation of the forces using the Peyrard-Bishop model (32, 57). According to this model the potential energy of a sequence of N + 1 base pairs whose first base pair (n = 0) is pulled by a force F perpendicular to the sequence is (57): where yn is the deviation from equilibrium of the distance between the bases of the n−th base pair. The first term describes the pairing interaction and the second one the stacking interaction. By (yn − yn+1)2] − F y0 K 2 (7) N(cid:88) n=0 V = [D(1 − e−ayn) + −40−35−30−25−20−15−10−5 0 0 0.5 1 1.5 2−TS [kcal/mol]Separation Distance[Å]unzip2nd bp 0 1 2 3 4 5 6 7 0 2 4 6 8 10 12 14Free Energy[kcal/mol]Separation Distance[A] 0 0.02 0.04 0.06 0.08 0.1 0.12 0 0.5 1 1.5 2Entropy[kcal/mol/K]Separation Distance[A] 0 1 2 3 4 5 6 7 0 2 4 6 8 10 12 14Free Energy[kcal/mol]Separation Distance[A] 0 0.02 0.04 0.06 0.08 0.1 0.12 0 0.5 1 1.5 2Entropy[kcal/mol/K]Separation Distance[A]unzip1st bp 0 1 2 3 4 5 6 7 0 2 4 6 8 10 12 14Free Energy[kcal/mol]Separation Distance [Å] 100% paired 'first' bp100% frayed 'first' bp10% frayed(cid:239)90% paired 'first' bp30% frayed(cid:239)70% paired 'first' bp(cid:239)40(cid:239)35(cid:239)30(cid:239)25(cid:239)20(cid:239)15(cid:239)10(cid:239)5 0 0 0.5 1 1.5 2(cid:239)TS [kcal/mol]Separation Distance[Å]o Initiation of DNA mechanical unzipping differentiating Eq. 7 with respect to yn, one obtains: = 2aDe−ay0(1 − e−ay0) + K(y0 − y1) − F ∂V ∂y0 = 2aDe−ayn(1 − e−ayn) + K(2yn − yn+1 − yn−1) ∂V ∂yn for n > 0 10 (8) Imposing that the conformation with minimum energy satisfies ∂V /∂yn = 0 for all values of n in the second equality of Eq. (8) and taking its continuum n limit, one gets d2u dn2 − Ae−u(1 − e−u) = 0 where u = ay and A = 2a2 D K . The two general solutions of this equation read − 1)e∓C1(n+C2)] e±C1(n+C2) + + u = ln[ A C2 1 1 4C1 A C1 ( A C2 1 (9) (10) where C1 and C2 are two integration constants. The physical solution requires that u tends towards 0 when n tends towards +∞, meaning that the far end of the sequence is still zipped, which is possible if and only if C2 1 = A. The physical solution can therefore be recast in the form: u(nthresh n) = ln[1 + e−√ √ 1 A 4 A(n−nthresh)] (11) where, due to the large value of A (see below), nthresh represents approximately the rank of the base pair up to which the sequence is unzipped. The first line of Eq. 8 is then used, together with the minimum condition ∂V /∂yn = 0, to determine the force F (nthresh) that is necessary to keep the sequence unzipped up to base pair nthresh, F (nthresh) = 2aDe−u(nthresh0)(1 − e−u(nthresh0)) + K a (u(nthresh 0) − u(nthresh 1)). (12) When substituting in the above equation the typical values for the parameters of the Peyrard- Bishop model (D = 0.063eV, K = 0.025eV A−2, and a = 4.2 A−1 (57) - so that A = 88.9056) we find that there is a high force barrier for opening the first base pair, see Fig. 3 (as discussed below, we also observe this barrier in mesoscopic and atomistic simulations). The position and height of the barrier can, in principle, be obtained analytically as a function of D, K and by searching for the maximum of F in Eq. 12, but the final expressions are too long and tedious to be reproduced here. Numerically, the force threshold is 218.69 pN. The validity of Eq. 12 was checked by integrating numerically Hamilton's equations of motion. This led to a force threshold within 1 pN of the value derived from Eq. 12. Initiation of DNA mechanical unzipping 11 Figure 3: Mean force on the terminal base pairs need to maintain them at a given separation, as a function of that separation distance. Atomistic simulations in red for fully paired first base-pair in red, in orange dashed line for the case of first base pair 10% frayed and 90%, light brown dashed for 30% frayed and 70% paired; together with the results of Ref. (37) in green dashed line, the Peyrard-Bishop model (dotted dark blue) and the coarse-grained simulations (pink dotted line). We also show (in light blue and black) results for coarse-grained simulations of the same sequence with the first base pair and second base pairs changed from C-G to A-T. The vertical lines mark the distances beyond which the first and second base pairs are fully unzipped. The force-separation curves obtained through the three methods are displayed in Fig. 3. The results obtained using the Peyrard-Bishop model are plotted with a blue dotted line, those obtained from all atom simulations with a red line for fully paired first base-pair, the orange dashed line atomistic simulations for 10% frayed and 90% paired first base-pair, the light brown dashed line atomistic simulations for 30% frayed and 70% paired first base-pair, and those of the coarse-grained simulations with a magenta dotted line. Moreover, the figure also displays results of coarse-grained simulations for two variations of the DNA sequence: in the first one, the first base pair has been changed from CG to AT (black dotted line), while in the second one the same substitution has been done for the second base pair (magenta dotted line). We note the presence not only of the initial high-force peak to separate the first base pair, but also the presence of an echo, a second peak of a relatively smaller force (appprox. 50 pN, but still larger than the "bulk" separation force). Finally, we also show, for comparison, the result of Ref. (37) as a green dashed line. In this work the authors develop a semi-microscopic model for the binding of the two nucleic acid strands which also predicts the presence of a high energetic barrier for DNA mechanical unzipping and accounts its origin to the higher rigidity of the double helix as compared to the single DNA strand. In the coarse-grained and all-atoms simulations, two force peaks are also observed at small separations: the first one is very sharp and occurs at the beginning of unzipping. Its magnitude (cid:18)(cid:84)(cid:85)(cid:1)(cid:35)(cid:66)(cid:84)(cid:70)(cid:1)(cid:49)(cid:66)(cid:74)(cid:83)(cid:19)(cid:79)(cid:69)(cid:1)(cid:35)(cid:66)(cid:84)(cid:70)(cid:1)(cid:49)(cid:66)(cid:74)(cid:83)o 0 50 100 150 200 250 300 0 2 4 6 8 10 12 14Force [pN]Separation Distance [A]atomistic(CGC...)-100% pairedatomistic(CGC...)-10% frayed-90% pairedatomistic(CGC...)-30% frayed-70% pairedcoarse-grained/CG (CGC....)coarse-grained/CG (TGC....)coarse-grained/CG (CTC....)PB ModelCocco et.al. Initiation of DNA mechanical unzipping 12 varies between 132 and 219 pN depending on the model used. This range is in agreement with the values predicted by other theoretical work (37, 39). It is noteworthy that, in the coarse-grained simulations displayed in Fig. 3, the second peak for for the two dodecamers that have CG base pairing at position 2 the peaks are identical, while for the dodecamer that has AT base pairing at the second position the peak is lower and occurs at slightly smaller interstrand displacement. Also within reason is the fact that when the first unzipped base pair is the same (CG) in two different dodecamers (i.e., in which the first basepair stacks on and AT vs a CG base pair) in the coarse grained simulations, the force profiles are nearly identical for the first peak (see Fig. 3). The observation of these two force peaks is explained by the free energy plot in Fig. 2, which has two local minima (with steeper slopes, yielding the force peaks) at the same separation distances at which the force peaks are observed. The inset to Fig. 2 shows that the conformational entropy part of the total entropy contribution to the free energy well around the first base pair for the fully paired first base-pair (black line in Fig. 2)) has a strong contribution, which accounts in part for the high force peak. These small inter-strand separations (certainly for the first peak and most likely also for the second one) are likely below the minimum resolution one can use to investigate them through typical AFM experiments (15 -- 17). Such peaks are yet to be observed experimentally in singe molecule experiments as increased force-distance resolutions become available. As an encouraging alternative, unzipping experiments have already recorded pausing events whose magnitude may well be associated with the overcoming of these energy barriers (11). An interesting feature is the second force-peak, located at the larger separation associated with the second base-pair rupture. It is weaker (∼ 50 pN) than the first peak and wider, and reminiscent of the unzipping "echoes" in Ref. (11). For even larger separations (echo diminishing, see above)), the force tends towards a constant value (the so called "critical force," in the large scale unzipping studies), which is the force needed to keep the two DNA strands separated. It has been shown experimentally (13) that this force is constant for homogenous sequences and fluctuates if the sequences are inhomogeneous. Our values for these long-scale separation forces are close to 20 pN, which is within the range of measured values (11, 14). Of note is also the fact that, while both the semi-microscopic model of Ref. (37), as well as the analytical PB treatments both lead to the appearance of a first peak, neither feature a second peak, which is indicative of the fact that they are, in effect, local models. The second peak is observed in both our atomistic and coarse-grained studies because they involve longer range interactions encompassing more degrees of freedom, hinting at a more nuanced picture for the balance of forces at play at the end of DNA duplexes. The fact that this high barrier for initiation of unzipping is observed in all descriptions gives, on one hand, information about its origin, but it is also a validation of the coarse-grained models. Moreover, by observing the gradual increase in the number of interactions included in these models, we can assess which are the main contributors to the observed force peaks. The Peyrard-Bishop model contains only two terms, namely stacking between consecutive bases on the same strand and pairing between complementary bases on opposite strands. Therefore, the high force needed to initiate DNA unzipping has to stem from to the need to overcome these two types of interactions in a manner that depends on whether the terminal or bulk base pairs are broken. This hypothesis is also confirmed by the fact that, when changing a base pair from CG to AT the magnitude of this force decreases accordingly (a GC base pair contains three hydrogen bonds, while an AT has only two, making it easier to break). Also in common, both coarse grained and atomistic models showed that the base pairing opening pathway was towards the major groove (see Movie in Supplemental Information), which is in accord with previous studies showing this direction as being more favorable (58 -- 60). In order to further check our hypothesis, we plotted separately some of the energy terms for Initiation of DNA mechanical unzipping 13 both the all atom and coarse-grained models as a function of the separation distance. They are displayed in Figs. 4 and 5 for the all atom simulations and the coarse-grained model, respectively. The first panel depicts the base pairing energy. Our coarse-grained model does not account for non- native pairing so we only plot the energy of the first base pair in red and that of the second one in dotted green. It can be seen clearly that there is a sharp increase in energy at a separation distance corresponding to the first and second force peaks, respectively. Moreover, the slope corresponding to the separation of the second base pair is weaker, thus accounting for the smaller magnitude of the second peak. This is confirmed by the hydrogen bonding energy terms in the all-atom model, which display similar tendencies (see also distance dependence of hydrogen bonding in Supplementary Fig. 1). For the all-atom simulations we also plot some of the non-native pairing terms. As expected, these terms display only small variations, suggesting that their contribution to the force peak is minor. The other panels of Figs. 4 and 5 display the van der Waals and electrostatic terms, separately (Fig. 4-b, c) and stacking (Fig. 5-b) and electrostatic (Fig. 5-c) interactions for the coarse-grained model, for several combinations of bases, either on the same strand or on different strands. Initiation of DNA mechanical unzipping 14 Figure 4: Select terms of the all-atom Hamiltonian as a function of separation distance, averaged over all simulation windows. a) Native hydrogen bonds energy between the first and second base pairs (the solid red and dashed green lines, respectively), and some non-native hydrogen bonds energies. b) van der Waals term between the first base pair and different bases on the same DNA strand and/or complementary DNA strands. c) Electrostatic interactions between the first base pair and different bases on the same DNA strand and/or complementary DNA strands. G12-G2C11-C3G2-C1G12-C3C1-G12G2-C11G12-C11G2-C1G12-C3G12-G2C11-C3C1-G12G2-C11C11-C3G12-C3G12-G2G12-C11G2-C1G12-C11C1-G12G2-C11(cid:233)18(cid:233)16(cid:233)14(cid:233)12(cid:233)10(cid:233)8(cid:233)6(cid:233)4(cid:233)2 0 0 2 4 6 8 10 12 14Hbond Energy[kcal/mol](cid:233)3(cid:233)2.5(cid:233)2(cid:233)1.5(cid:233)1(cid:233)0.5 0 0.5 1 1.5 2 0 2 4 6 8 10 12 14VDW Energy[kcal/mol](cid:233)35(cid:233)30(cid:233)25(cid:233)20(cid:233)15(cid:233)10(cid:233)5 0 5 10 15 20 0 2 4 6 8 10 12 14Elec. Energy[kcal/mol]a)b)c)-20-10 0 10 20 30 40 50 0 2 4 6 8 10 12 14Force [pN]x-directiony-directionz-direction-40-20 0 20 40 0 2 4 6 8 10 12 14Force [pN]x-directiony-directionz-direction-100 0 100 200 300 400 500 600 700 0 2 4 6 8 10 12 14Torque [pN.A]Separation Distance [A]torque totaltorque total in z-dir -100 0 100 200 300 400 500 600 700 0 2 4 6 8 10 12 14Torque [pN.A]Separation Distance [A]torque totaltorque total in z-dir -100 0 100 200 300 400 500 600 700 0 2 4 6 8 10 12 14Torque [pN.A]Separation Distance [A]torque totaltorque total in z-dir ooG2-C1C11-C3G12-C11C1-G12G12-C3a)G2-C11G12-G2(cid:239)18(cid:239)16(cid:239)14(cid:239)12(cid:239)10(cid:239)8(cid:239)6(cid:239)4(cid:239)2 0 0 2 4 6 8 10 12 14Hbond Energy[kcal/mol]Separation Distance(cid:239)[Å]ob)C1-G12G2-C1G12-C11G12-G2C11-C3G12-C3G2-C11(cid:239)3(cid:239)2.5(cid:239)2(cid:239)1.5(cid:239)1(cid:239)0.5 0 0.5 1 1.5 2 0 2 4 6 8 10 12 14VDW Energy[kcal/mol]Separation Distance(cid:239)[Å]ob)a)c)C1-G12C11-C3G2-C1G12-C11G12-C3G12-G2G2-C11(cid:239)35(cid:239)30(cid:239)25(cid:239)20(cid:239)15(cid:239)10(cid:239)5 0 5 10 15 20 0 2 4 6 8 10 12 14Elec. Energy[kcal/mol]Separation Distance(cid:239)[Å]oc)c)C1-G12C11-C3G2-C1G12-C11G12-C3G12-G2G2-C11(cid:239)35(cid:239)30(cid:239)25(cid:239)20(cid:239)15(cid:239)10(cid:239)5 0 5 10 15 20 0 2 4 6 8 10 12 14Elec. Energy[kcal/mol]Separation Distance(cid:239)[Å]o Initiation of DNA mechanical unzipping 15 Figure 5: Select terms of the coarse-grained Hamiltonian as a function of separation distance, averaged over all simulation runs. a) Pairing energies between the first and second base pairs (the solid red line and the green dots, respectively).b) The base stacking energies between first and second and first and third bases on one of the two DNA strand (the curves for the second strand are similar) c) Electrostatic interactions between the first base pair and different bases on the complementary DNA strands. At this point, comparison between the two types of representations becomes more tedious, because of the simplifications made in the construction of the coarse-grained model. Firstly, coarse- grained electrostatic interactions only act between phosphates and only have a repulsive part, as can be seen in Fig. 5c. They essentially contribute to prevent different DNA segments from overlapping, but they also increase the rigidity and the persistence length of the backbones. The stacking interactions could be compared to the sum of the van der Waals and electrostatic terms for the CHARMM force field. Both representations predict that the variation of these energy terms Initiation of DNA mechanical unzipping 16 is rather small compared to that of the pairing/hydrogen bonding terms. Nevertheless, there is a stacking barrier at the separation distances where the two peaks occur (Fig. 4b), which shows that they do contribute to the high force needed to initiate unzipping. This is also seen from the intermediate atomic structures displayed in Fig. 6, which depicts six snapshots of some of the conformations that DNA takes during unzipping, as extracted from the umbrella sampling MD simulations and plotted using VMD. The first panel shows the equilibrated sequence and the red dots show the points where the separating force is applied. The second conformation corresponds to roughly the same separation distance where the force peak occurs. Note an intra-strand "bond" between the first two bases on the second strand (G12-C11). This is also confirmed by the plot of the energy (van der Waals term) in Fig. 4-b and the stacking energy in Fig. 5-b. This bond is broken at larger separations (the second configuration), where a transient across-strand bond between G12 and C3 (i.e., a "non-native" H-bond) is formed, and then again transiently reforms. Moreover, this also happens for several other non-native H-bonds (as seen in the fourth and sixth panels of Fig. 6). In the fifth and sixth panels, the second base pair is already opening, but now intra-strands bonds between bases seem to have recovered, as also suggested by the van der Waals G2-C1 and G12-C11 energy terms plotted in Fig. 4-b. More generally, we observe that these non- native interactions have either a "hydrogen bond" or electrostatic character, rather than a van der Waals interaction. It is however difficult to evaluate how much these changes in the interactions along one strand could ultimately contribute to a diminution of its rigidity. A similar trend is observed for the plots of the stacking energy in the coarse-grained simulations (see Fig. 5-b, where the energy terms are only shown for one DNA strand, but are similar for the other one): the energy for the G12-C11 interaction increases at higher separations and then decreases again. It is fair to admit that our coarse-grained model is not detailed enough to capture the change in the nature of the forces determining the interactions. However, it can be observed from visual inspection of the atomic structures in Fig. 6 (see also Movie in SI), and also from coarse-grained calculations (data not shown) that some of the bases rotate when opening, as suggested by the simulations of reference (39). That work also suggested the presence of a torsional barrier to unzipping, which is confirmed in our atomistic simulations (see Supplementary Material). Initiation of DNA mechanical unzipping 17 Figure 6: Snapshots of some of the intermediate structures that form while unzipping DNA, ex- tracted from our atomistic simulations using VMD. Transient bonds that form during unzipping are shown with a dashed blue line. Nucleotides G, C, A and T are given in colors yellow, orange, blue and magenta, respectively. The two DNA strand backbones are shown in blue and in red. Taken together, our results suggest that the occurrence of a force barrier when opening the first DNA base pair has two main causes: the breaking of hydrogen bonds between the first bases on each strand, and that of stacking interactions between these bases and their nearest-neighbor along the same strand. When the sequence is completely zipped, there are few fluctuations in its conformation and this is seen in the high forces needed to break the hydrogen bonds. Once the bonds forming the first base pair are broken, the bases have access to more configurations (as evidenced by the higher entropy, see inset to Fig. 2), fluctuations increase (the chain is also more flexible) and the second base pairs are easier to open. This phenomenology can also be used to describe the microscopic origin of the cooperativity manifested by statistical models. It also agrees with previous simulations of DNA base pair opening which suggest that a strictly local model of the opening of DNA base pairs would not hold (59, 60). An additional factor to consider is the contribution to the force from the solvent, which was shown to be a major determinant for the dsDNA helical conformation (38). Moreover, the interpretation of the initial barrier as having an important contribution from stacking interactions within the same strand (in addition to the breaking of hydrogen bonds) is in line with earlier calculations on the cost of unstacking (61, 62), which showed a 2-4 kcal/mol barrier before the bases become independently solvated at about 2A increased separation The presence of such a high force barrier involved in DNA mechanical unzipping has already been discussed in the literature, but the presence of a second peak has, to our knowledge, not been previously reported. This second peak is especially well observed in the atomistic simulations, because of their high resolution. We moreover expect that, if these simulations would be run DNAs1DNAs2at 0 Angstrom 3`5`G12:O3`C1:O3`15.10DNAs1DNAs2at 2.30 Angstrom 3`5`Non-native H-BondG12:O3`C1:O3`17.44C3:O3`3`5`at 3.90 Angstrom DNAs2DNAs1G12:O3`C1:O3`19.13G2:O3`C11:O3`15.86DNAs1DNAs25`3`at 9.90 Angstrom G12:O3`C1:O3`25.15C11:O3`DNAs1DNAs23`5`at 12.90 Angstrom G12:O3`C1:O3`28.42G2:O3`C11:O3`24.52DNAs13`DNAs25`at 13.15 Angstrom Non-native H-BondG12:O3`C1:O3`28.58C11:O3` Initiation of DNA mechanical unzipping 18 for a higher separation and for a longer sequence, a third peak corresponding to the opening of the third base pair could be observed. Although our coarse-grained simulations were run until the separation distance has reached 50 A we do not observe other force peaks because at higher separations denaturation will proceed in a very fast and irregular manner, with significant noise. This is confirmed by the plot of the number of open base pairs as a function of separation: it shows a large increase at distances higher that 20 A, similar to the phase transition that has been observed for larger sequences (34). We however ran a set of coarse-grained simulations on a homogeneous CG sequence of similar length, and there we do observe the occurrence of a third peak at a separation distance of ∼ 18A, which is smaller and wider than the second one (data not shown). Our findings are supported by an earlier atomistic simulation study of DNA mechanical denaturation using the AMBER force field (39). There, the authors applied an increasing force perpendicular to the helix axis, and plotted the separation and the number of open base pairs as a function of the force, for various temperatures. For 300 K, which is closest to the temperature we use, they observe that the base pairs start to separate at a critical force of 237 pN, after which unzipping progresses through jumps and pauses. We note, in passing, that this involves and out-of-equilibrium situation, and the use of a different simulation protocol, which does not use the umbrella sampling restraints in our simulations. Similar discrepancies between the experimental force and the rapidly pulling force in simulations was previously discussed in studies of the end to end stretching of DNA (63). As discussed in the Introduction, experimental studies have established the fact that DNA ends fray in solution - that is, that the terminal base pair can open spontaneously due to thermal fluctuations. For a sequence with a GC terminal base pair, the fraying probability is around 10%. We accounted for this in our work by running an additional set of simulations in which the first base pair is initially open (frayed) and then computing a weighted averaged force profile between the two configurations (paired and frayed) with weights that correspond to the experimental probabilities of the open and close state. We also point out that the requirement in statistical mechanical models for a large force to initiate unzipping is not in conflict with the observations of DNA terminal base- pair fraying. For example, using Langer barrier-crosssing theory, Cocco et al (37) have shown that there occurs a fluctuation-assisted crossing of the free energy barrier for opening corresponding to this force. It is instructive at this point to make a comparison between the resolutions and accuracies of the two types of simulations used in this work. The fact that they both manage to capture the position of the two force peaks proves, on one hand, the robustness of the coarse-grained model and, on the other, the entropic-well origin derived from the atomistic representation. The magnitude of the critical force is different in all descriptions used herein, and its value actually decreases when the resolution of the model is increased, suggesting that the coarse-grained models would tend to overestimate this force. One would have expected an opposite trend, since in the simpler model there is a smaller number of configurations available to the system, and thus reduced entropy. On the other hand, in the all atom case one captures more intermediate states with energy very close to that of the fully unzipped base pair, making the transition less abrupt. Although both types of models manage to describe the main phenomena at similar accuracy, all atom simulations bring more information on the details of the intermediate states that occur during unzipping and on their dynamics. Some aspects, like the various types of interactions that form between neighboring bases along the same strand or non-Watson-Crick inter-strand H-bonds can only be captured by atomistic simulations. An ideal method would be a combination of the two: the coarse-grained models would allow the simulations of larger durations, and the more interesting events could then be simulated in more details using atomistic force fields. Initiation of DNA mechanical unzipping 19 4 Conclusion We used molecular simulations at two resolution levels and an analytical (Peyrard-Bishop) model to provide a detailed study of the onset of DNA mechanical denaturation. Our results bring new information on the transient interactions that occur during this process. We observe a large force peak at ∼ 2 A separation and a second, smaller peak at distances ranging between 8 and 12 A. We predict that the force peaks on the profile will continue, but with lower values, for the opening of the following base pairs (a third small peak is seen in the coarse-grained simulations, data not shown) but will become indiscernible due to an increase in the "signal-to-noise" ratio. To understand the origin of these force peaks, we have computed free energy profiles and we have further analyzed conformational entropy contributions, hydrogen bonding interactions (for both native, i.e., canonical Watson-Crick pairing, and for non-native connections) and stacking interactions of the first few bases at the end of the DNA molecule where the unzipping force is applied. We observe secondary contributions to the force peaks from entropic effects associated with the other types of interactions within the DNA sequence. The essential feature leading to the presence of the initial large force peak(s) comes from analyzing the potential of mean force. The first well is narrower than the second and the subsequent ones because of less entropy (fewer states in the reaction coordinate), hence it is steeper, leading to a larger slope. The force needed for unzipping is thus higher because of the lower entropy of the chains zipped up for the first base pair as opposed to the second one, and this effect diminishes in "ripples," or echoes, as the separation between the strands increases. Eventually, when a significant portion of DNA is in single-stranded form, separation forces drop to the 12-20 pN limit observed in the experiments, as base-pair separation becomes akin to melting, which is known to be driven by fluctuations and therefore strongly depends on the conformations available to the now-floppier single-stranded force handles. The observed higher forces for unzipping initiation relative to the forces needed in single molecule pulling point to a difference in behavior in boundary vs. bulk pairs. While to date we are not aware of any direct experimental probing to confirm or disapprove the presence of the large forces needed for initiation, indirect confirmation may exist. For example, proton exchange has been used to probe base-pair opening kinetics in 5'-d(CGCGAATTCGCG)-3' and related dodecamers (64). The enthalpy changes for opening of the central basepairs are correlated to the opening entropy changes. This enthalpy-entropy compensation minimizes the variations in the opening free energies among these central basepairs. Deviations from the enthalpy-entropy compensation pattern are observed for basepairs located close to the ends of the duplex structure, suggesting a different mode of opening for these basepairs. It is possible that the difference in unzipping the first base pairs revealed in our work could be a manifestation of this difference in opening modes observed in the NMR data. Longer simulations at the actual separation forces with a variety of atomistic DNA force fields or more sophisticated sampling schemes of the actual kinetics of the transition (65), together with new experimental techniques developing increasingly in resolution (such as force clamp spectroscopy (66) or nanopore unzipping technologies (2)) may give more insight into the sequence dependent thermodynamics and kinetics of Watson-Crick (67) and alternative (68) base-pairing phenomena. Author Contributions A.M., A.M.F. and I.A. designed research; A.M., A.M.F., E.B. and J.W. performed research; A.M.F., M.J. and I.A. contributed analytical tools; A.M., A.M.F. and I.A. analyzed data; A.M., A.M.F., Initiation of DNA mechanical unzipping 20 E.B., I.A. wrote the paper. Acknowledgments IA acknowledges support from NIH (grant 5R01GM089846) and the NSF (grant CMMI-0941470). A. M. F. thanks the MPG for support by a postdoctoral grant through the MPG-CNRS GDRE "Systems Biology." We also acknowledge NSF Grant CHE-0840513 for the computational resources used for our calculations on the Greenplanet cluster at UC Irvine. Supplementary Information Several other analyses were performed, gauging the force components and the torque on the terminal base pairs upon unzipping and an additional hydrogen bonding analyses; figures and discussion are included as Supplementary Material. We also uploaded a Supplementary Movie showing the unzipping of the first two base-pairs from the umbrella sampling trajectories. The movie was made taking 20 frames from each umbrella sampling window and used VMD. Nucleotides G, C, A and T are given in colors yellow, orange, blue and magenta, respectively. The two DNA strand backbones are shown in blue and in red. The C1:O3(cid:48) and G12:O3(cid:48) atoms that are pulled apart are shown as red balls and their distance is marked with black dotted lines. The first two base-pairs are in opaque and the rest of DNA in transparent representations. References 1. Saenger, W., 1984. Principles of nucleic acid structure. Springer-Verlag, New York. 2. McNally, B., M. Wanunu, and A. Meller, 2008. Electromechanical Unzipping of Individual DNA Molecules Using Synthetic Sub-2 nm Pores. Nano Lett. 8:3418 -- 3422. 3. Boland, T., and B. D. Ratner, 1995. Direct measurement of hydrogen bonding in DNA nu- cleotide bases by atomic force microscopy. Proc. Natl. Acad. Sci. U.S.A. 92:5297 -- 5301. 4. Bustamante, C., S. B. Smith, J. Liphardt, and D. Smith, 2000. Single-molecule studies of DNA mechanics. Curr. Opin. Struct. Biol. 10:279 -- 285. 5. Strick, T. R., M. N. Dessinges, G. Charvin, N. H. Dekker, J. F. Allemand, D. Bensimon, and V. Croquette, 2003. Stretching of macromolecules and proteins. Rep. Prog. Phys. 66:1 -- 45. 6. Meller, A., 2003. Dynamics of polynucleotide transport through nanometre-scale pores. J. Phys.- Condens. Matter 15:R581 -- R607. 7. Hirsh, A. D., M. Taranova, T. A. Lionberger, T. D. Lillian, I. Andricioaei, and N. C. Perkins, 2013. Structural Ensemble and Dynamics of Toroidal-like DNA Shapes in Bacteriophage phi 29 Exit Cavity. Biophys. J. 104:2058 -- 2067. 8. Mao, C., W. Sun, Z. Shen, and N. C. Seeman, 1999. A nanomechanical device based on the B-Z transition of DNA. Nature 397:144 -- 146. 9. Seeman, N. C., 1999. DNA engineering and its application to nanotechnology. Trends in Biotechnology 17:437 -- 443. Initiation of DNA mechanical unzipping 21 10. Danilowicz, C., Y. Kafri, R. S. Conroy, V. W. Coljee, J. Weeks, and M. Prentiss, 2004. Mea- surement of the phase diagram of DNA unzipping in the temperature-force plane. Phys. Rev. Lett. 93:078101 -- 4. 11. Danilowicz, C., V. W. Coljee, C. Bouzigues, D. K. Lubensky, D. R. Nelson, and M. Prentiss, 2003. DNA unzipped under a constant force exhibits multiple metastable intermediates. Proc. Natl. Acad. Sci. U.S.A. 100:1694 -- 1699. 12. Weeks, J. D., J. B. Lucks, Y. Kafri, C. Danilowicz, D. R. Nelson, and M. Prentiss, 2005. Pause Point Spectra in DNA Constant-Force Unzipping. Biophys. J. 88:2752 -- 2765. 13. Essevaz-Roulet, B., U. B. F., and Heslot, 1997. Mechanical separation of the complementary strands of DNA. Proc. Natl. Acad. Sci. U.S.A. 94:11935 -- 11940. 14. Rief, M., H. Clausen-Schaumann, and H. E. Gaub, 1999. Sequence dependent mechanics of single DNA molecules. Nat. Struct. Biol. 6:346 -- 349. 15. Albrecht, C., K. Blank, M. Lalic-Multhaler, S. Hirler, T. Mai, I. Gilbert, S. Schiffmann, T. Bayer, H. Clausen-Schaumann, and H. E. Gaub, 2003. DNA: A programmable force sensor. Science 301:367 -- 370. 16. Strunz, T., K. Oroszlan, R. Schafer, and H. J. Guntherodt, 1999. Dynamic force spectroscopy of single DNA molecules. Proc. Natl. Acad. Sci. U.S.A. 96:11277 -- 11282. 17. Krautbauer, R., M. Rief, and H. E. Gaub, 2003. Unzipping DNA oligomers. Nano Lett. 3:493 -- 496. 18. Wong, K. Y., and B. M. Pettitt, 2008. The pathway of oligomeric DNA melting investigated by molecular dynamics simulations. Biopys. J. 95 (12):5618 -- 5626. 19. Zgarbova, M., M. Otyepka, J. Sponer, F. Lankas, and P. Jurecka, 2014. Base Pair Fraying in Molecular Dynamics Simulations of DNA and RNA. J. of Chem. Theory Comput. 10:3177 -- 3189. 20. Kornberg, A., and T. Baker, 2005. DNA Replication. University Science Books. 21. Katz, R. A., G. Merkel, M. D. Andrake, H. Roder, and A. M. Skalka, 2011. Retroviral integrases promote fraying of viral DNA ends. J. Biol. Chem. 286 (29):25710 -- 25718. 22. Andricioaei, I., A. Goel, D. R. Herschbach, and M. Karplus, 2004. Dependence of DNA Poly- merase Replication Rate on External Forces: A Model Based on Molecular Dynamics Simula- tions. Biophysical Journal 87:1478 -- 1497. 23. Nonin, S., J. L. Leroy, and M. Gueron, 1995. Terminal base pairs of oligodeoxynucleotides: imino proton exchange and fraying. Biochemistry 34 (33):10652 -- 10659. 24. Kochoyan, M., G. Lancelot, and J. L. Leroy, 1998. Study of structure, base-pair opening kinetics and proton exchange mechanism of the d-(AATTGCAATT) self-complementary oligodeoxynu- cleotide in solution. Nucleic Acids. Res. 16 (15):7685 -- 7702. 25. Leroy, J. L., M. Kochoyan, T. Huynh-Dinh, and M. Gueron, 1988. Characterization of base-pair opening in deoxynucleotide duplexes using catalyzed exchange of the imino proton. J. Mol. Biol. 200 (2):223 -- 238. Initiation of DNA mechanical unzipping 22 26. Jafilan, S., L. Klein, C. Hyun, and J. Florin, 2012. Intramolecular Base Stacking of Dinucleoside Monophosphate Anions in Aqueous Solution. The Journal of Physical Chemistry B 116:3613 -- 3618. 27. Sebastian, K. L., 2000. Pulling a polymer out of a potential well and the mechanical unzipping of DNA. Phys. Rev. E 62:1128 -- 1132. 28. Bhattacharjee, S. M., 2000. Unzipping DNAs: towards the first step of replication. J. of Phys. A 33:23 -- 28. 29. Kafri, Y., D. Mukamel, and L. Peliti, 2002. Denaturation and unzipping of DNA: statistical mechanics of interacting loops. Physica A 306:39 -- 50. 30. Lubensky, D. K., and D. R. Nelson, 2000. Pulling pinned polymers and unzipping DNA. Phys. Rev. Lett. 85:1572 -- 1575. 31. Marenduzzo, D., S. M. Bhattacharjee, A. Maritan, E. Orlondini, and F. Seno, 2001. Dynamical scaling of the DNA unzipping transition. Phys. Rev. Lett. 88:028102 1 -- 4. 32. Peyrard, M., and A. R. Bishop, 1989. Statistical-mechanics of a nonlinear model for DNA denaturation. Phys. Rev. Lett. 62:2755 -- 2758. 33. Dauxois, T., M. Peyrard, and A. Bishop, 1993. Entropy driven DNA denaturation. Phys. Rev. E. 47:R44 -- R48. 34. Peyrard, M., 2004. Nonlinear dynamics and statistical physics of DNA. Nonlinearity 2004:R1 -- R40. 35. Zdravkovic, S., and M. V. Sataric, 2006. Single-molecule unzippering experiments on DNA and Peyrard-Bishop-Dauxois model. Phys. Rev. E 73:021905 1 -- 11. 36. Cocco, S., R. M. J. F., and Marko, 2001. Force and kinetic barriers to unzipping of the DNA double helix. Proc. Natl. Acad. Sci. U.S.A. 98:8608 -- 8613. 37. Cocco, S., R. Monasson, and J. F. Marko, 2002. Force and kinetic barriers to initiation of DNA unzipping. Phys. Rev. E 65:041907, 1 -- 23. 38. Maffeo, C., J. Yoo, J. Comer, D. B. Wells, B. Luan, and A. Aksimentiev, 2014. Close encounters with DNA. Journal of Physics: Condensed Matter 26:413101. 39. Santosh, M., and P. K. Maiti, 2009. Force induced DNA melting. J. Phys.-Condens. Mat. 21:034113 1 -- 8. 40. Florescu, A. M., and M. Joyeux, 2011. Thermal and mechanical denaturation properties of a DNA model with three sites per nucleotide. J. Chem. Phys. 135:085105 1 -- 12. 41. Brooks, B. R., C. L. Brooks, A. D. Mackerell, L. Nilsson, R. J. Petrella, B. Roux, Y. Won, G. Archontis, C. Bartels, S. Boresch, A. Caflisch, L. Caves, Q. Cui, A. R. Dinner, M. Feig, S. Fischer, J. Gao, M. Hodoscek, W. Im, K. Kuczera, T. Lazaridis, J. Ma, V. Ovchinnikov, E. Paci, R. W. Pastor, C. B. Post, J. Z. Pu, M. Schaefer, B. Tidor, R. M. Venable, H. L. Woodcock, X. Wu, W. Yang, D. M. York, and M. Karplus, 2009. CHARMM: The biomolecular simulation program. J. Comput. Chem. 30:1545 -- 1614. Initiation of DNA mechanical unzipping 23 42. Foloppe, N., and A. MacKerell, 2000. All-Atom Empirical Force Field for Nucleic Acids: 1) Parameter Optimization Based on Small Molecule and Condensed Phase Macromolecular Target Data. J. Comput. Chem. 21:86 -- 104. 43. MacKerell, A. D., and N. Banavali, 2000. All-Atom Empirical Force Field for Nucleic Acids: 2) Application to Molecular Dynamics Simulations of DNA and RNA in Solution. J. Comput. Chem. 21:105 -- 120. 44. Kumar, S., D. Bouzida, R. H. Swendsen, P. A. Kollman, and J. M. Rosenberg, 1992. The Weighted Histogram Analysis Method for free-energy calculations on biomolecules 1: The Method. J. Comput. Chem. 13:1011 -- 1021. 45. Grossfield, A. WHAM: the Weighted Histogram Analysis Method; version 2.0.6; http://membrane.urmc.rochester.edu/content/wham . 46. Jorgensen, W. L., J. Chandrasekhar, J. D. Madura, R. Impey, and M. L. Klein, 1983. Com- parison of simple potential functions for simulating liquid water. J. Chem. Phys . 79:926 -- 935. 47. Essmann, U., L. Perera, M. L. Berkowitz, T. Darden, H. Lee, and L. G. Pedersen, 1995. A Smooth Particle Mesh Ewald Method. J. Chem. Phys . 103:8577 -- 8593. 48. Nose, S., 1984. A unified formulation of the constant temperature molecular-dynamics methods. J. Chem. Phys . 81:511 -- 519. 49. Hoover, W. G., 1985. Canonical dynamics: Equilibrium phase-space distributions. Phys. Rev. A 31:1695 -- 1697. 50. Brooks, B. R., R. E. Bruccoleri, B. D. Olafson, D. J. States, S. Swaminathan, and M. Karplus., 1983. CHARMM: A program for macromolecular energy, minimization and dynamics calcula- tion. J. Comput. Chem. 4:187 -- 217. 51. Ryckaert, J.-P., G. Ciccotti, and H. Berendsen, 1977. Numerical Integration of the Cartesian Equations of Motion of a System with Constraints: Molecular Dynamics of n-Alkanes. J. Comput. Phys. 23:327 -- 341. 52. Kirkwood, J. G., 1935. Statistical mechanics of fluid mixtures. J. Chem. Phys. 3:300 -- 313. 53. Wereszczynski, J., and I. Andricioaei, 2006. On structural transitions, thermodynamic equi- librium, and the phase diagram of DNA and RNA duplexes under torque and tension. Proc. Natl. Acad. Sci. U.S.A. 103:16200 -- 16205. 54. Knotts, T. A., N. Rathore, D. C. Schwartz, and J. J. de Pablo, 2007. A coarse grain model for DNA. J. Chem. Phys. 126:084901 1 -- 12. 55. Karplus, M., and J. N. Kushick, 1981. Method for estimating the configurational entropy of macromolecules. Macromolecules 14:325 -- 332. 56. Andricioaei, I., and M. Karplus, 2001. On the calculation of entropy from covariance matrices of the atomic fluctuations. J. Chem. Phys. 115:6289 -- 6292. 57. Singh, N., and Y. Singh, 2005. Statistical theory of force-induced unzipping of DNA. Eur. Phys. J. E 17:7 -- 19. Initiation of DNA mechanical unzipping 24 58. Ramstein, J., and R. Lavery, 1988. Energetic coupling between DNA bending and base pair opening. Proc. Natl. Acad. Sci. U.S.A. 85:7231 -- 7235. 59. Banavali, N. K., and A. D. MacKerell, 2002. Free Energy and Structural Pathways of Base Flipping in a DNA-GCGC Containing Sequence. J. Mol. Biol. 319:141 -- 160. 60. Giudice, E., P. V`arnai, and R. Lavery, 2003. Base pair opening within BDNA: free energy pathways for GC and AT pairs from umbrella sampling simulations. Nucleic Acids Res. 31:1434 -- 1443. 61. Norberg, J., and L. Nilsson, 1995. Potential of mean force calculations of the stacking- unstacking process in single-stranded deoxyribodinucleoside monophosphates. Biophys. J. 69:2277 -- 2285. 62. Norberg, J., and L. Nilsson, 1995. Stacking Free Energy Profiles for All 16 Natural Ribodinu- cleoside Monophosphates in Aqueous Solution. J. Am. Chem. Soc. 117:10832 -- 10840. 63. Harris, S. A., Z. A. Sands, and C. A. Laughton, 2005. Molecular dynamics simulations of duplex stretching reveal the importance of entropy in determining the biomechanical properties of DNA. Biophysical Journal 88:1684 -- 1691. 64. Moe, J., and I. Russu, 1990. Proton exchange and base-pair opening kinetics in 5'- d(CGCGAATTCGCG)-3' and related dodecamers. Nucleic Acids Research 18:821 -- 827. 65. Nummela, J., and I. Andricioaei, 2007. Exact low-force kinetics from high-force single-molecule unfolding events. Biophys. J. 93:3373 -- 3381. 66. Fernandez, J. M., and H. Li, 2004. Force-Clamp Spectroscopy Monitors the Folding Trajectory of a Single Protein. Science 303:1674 -- 1678. 67. Nikolova, E. N., G. D. Bascom, I. Andricioaei, and H. M. Al-Hashimi, 2012. Probing Sequence- Specific DNA Flexibility in A-Tracts and Pyrimidine-Purine Steps by Nuclear Magnetic Reso- nance C-13 Relaxation and Molecular Dynamics Simulations. Biochemistry 51(43):8654 -- 8664. 68. Nikolova, E. N., E. Kim, A. A. Wise, P. J. O'Brien, I. Andricioaei, and H. M. Al-Hashimi, 2011. Transient Hoogsteen base pairs in canonical duplex DNA. Nature 470:498 -- U84.
1707.09426
1
1707
2017-07-28T21:55:41
What dominates the time dependence of diffusion transverse to axons: Intra- or extra-axonal water?
[ "physics.bio-ph", "physics.med-ph", "q-bio.NC" ]
Brownian motion of water molecules provides an essential length scale, the diffusion length, commensurate with cell dimensions in biological tissues. Measuring the diffusion coefficient as a function of diffusion time makes in vivo diffusion MRI uniquely sensitive to the cellular features about three orders of magnitude below imaging resolution. However, there is a longstanding debate, regarding which contribution --- intra- or extra-cellular --- is more relevant in the overall time-dependence of the diffusion metrics. Here we resolve this debate in the human brain white matter. By varying not just the diffusion time, but also the gradient pulse duration of a standard diffusion pulse sequence, we identify a functional form of the measured time-dependent diffusion coefficient transverse to white matter tracts in 5 healthy volunteers. This specific functional form is shown to originate from the extra-axonal space, and provides estimates of the fiber packing correlation length for axons in a bundle. Our results offer a metric for the outer axonal diameter, a promising candidate marker for demyelination in neurodegenerative diseases. From the methodological perspective, our analysis demonstrates how competing models, which describe different physics yet interpolate standard measurements equally well, can be distinguished based on their prediction for an independent "orthogonal" measurement.
physics.bio-ph
physics
What dominates the time dependence of diffusion transverse to axons: Intra- or extra-axonal water? Hong-Hsi Leea,∗, Els Fieremansa, Dmitry S. Novikova aCenter for Biomedical Imaging, Department of Radiology, NYU School of Medicine, New York, NY 10016 Abstract Brownian motion of water molecules provides an essential length scale, the diffusion length, commensurate with cell dimensions in biological tissues. Measuring the diffusion coefficient as a function of diffusion time makes in vivo diffusion MRI uniquely sensitive to the cellular features about three orders of magnitude below imaging resolution. However, there is a longstanding debate, regarding which contribution -- intra- or extra-cellular -- is more relevant in the overall time-dependence of the diffusion metrics. Here we resolve this debate in the human brain white matter. By varying not just the diffusion time, but also the gradient pulse duration of a standard diffusion pulse sequence, we identify a functional form of the measured time-dependent diffusion coefficient transverse to white matter tracts in 5 healthy volunteers. This specific functional form is shown to originate from the extra-axonal space, and provides estimates of the fiber packing correlation length for axons in a bundle. Our results offer a metric for the outer axonal diameter, a promising candidate marker for demyelination in neurodegenerative diseases. From the methodological perspective, our analysis demonstrates how competing models, which describe different physics yet interpolate standard measurements equally well, can be distinguished based on their prediction for an independent "orthogonal" measurement. Keywords: diffusion, white matter, microstructure, time dependence, model selection 1. Introduction The ultimate promise of diffusion MRI (dMRI) (Jones, 2011), a technique that maps the diffusion propagator in each imag- ing voxel, is to become sensitive and specific to tissue features at the cellular level, orders of magnitude below the nominal imaging resolution. The foundation for this sensitivity is pro- vided by the diffusion length, i.e. the rms displacement of wa- ter molecules, being of the order of a few µm, which is com- mensurate with cellular dimensions. By controlling the diffu- sion time, one can probe the time-dependent diffusive dynamics (Tanner, 1979; Mitra et al., 1992; Assaf and Basser, 2005; As- saf et al., 2008; Alexander et al., 2010; Novikov et al., 2014; Burcaw et al., 2015; Fieremans et al., 2016; Reynaud et al., 2016), and quantify the relevant cellular-level tissue structure indirectly, using biophysical modeling (Yablonskiy and Suk- stanskii, 2010; Kiselev, 2017; Novikov et al., 2016a). In most tissues, and in the human brain in particular, the dMRI signal generally originates from at least two "compart- ments" -- intra- and extra-cellular spaces (Ackerman and Neil, 2010). Their distinct microgeometries provide different com- peting contributions to the overall non-Gaussian diffusion(Assaf and Basser, 2005; Alexander et al., 2010; Assaf et al., 2008; Fieremans et al., 2016; Burcaw et al., 2015). For any microstruc- tural interpretation of MRI experiments, it is crucial to deter- mine which contribution dominates, and which associated µm- level length scale can be in principle quantified. ∗Corresponding author Email address: [email protected] (Hong-Hsi Lee) Here we consider diffusion in human white matter (WM), transverse to major WM tracts. For the past decade, the fo- cus of microstructural modeling has been solely on the intra- axonal compartment, where the nontrivial (fully restricted) dif- fusion was thereby related to the inner axonal diameters (Assaf and Basser, 2005; Assaf et al., 2008; Alexander et al., 2010), whereas the extra-axonal diffusion has been deemed trivial (Gaus- sian). This framework has served as the basis for a number of techniques (CHARMED (Assaf and Basser, 2005), AxCaliber (Assaf et al., 2008), ActiveAx (Alexander et al., 2010)) for ax- onal diameter mapping. Their outcomes were subsequently de- bated due to a notable (Innocenti et al., 2015), sometimes by an order-of-magnitude (Alexander et al., 2010), overestimation of human inner axonal diameters relative to their histological values of ∼ 1 µm (Aboitiz et al., 1992; Caminiti et al., 2009; Liewald et al., 2014; Tang and Nyengaard, 1997; Tang et al., 1997). This recently prompted an alternative suggestion (Fiere- mans et al., 2016; Burcaw et al., 2015) of the dominant role of non-Gaussian, time-dependent diffusion in the extra-axonal space, with the role of the intra-axonal space deemed trivial (negligible radial signal attenuation due to thin axons). Rel- evant parameters for the extra-axonal picture characterize the packing geometry in a bundle; e.g., the packing correlation length should give a measure of outer axonal diameters (Fiere- mans et al., 2016; Burcaw et al., 2015). Since both alternatives have compelling arguments behind them and "fit the data well" (Alexander et al., 2010; Assaf et al., 2008; Fieremans et al., 2016; Burcaw et al., 2015), model selec- tion blindly based on fit quality is unreliable. This is a common 7 1 0 2 l u J 8 2 ] h p - o i b . s c i s y h p [ 1 v 6 2 4 9 0 . 7 0 7 1 : v i X r a challenge of model selection. To address it, here we (i) focus on the functional form of the competing models originating from their different physical assumptions, and (ii) use the fact that a true model would not just interpolate the standard measurement (varying the diffusion time) where both models perform well, but would also predict the outcome of an independent "orthog- onal" measurement. For the latter, we vary the gradient pulse width, which was not previously explored. Technically, we consider the dependence of the apparent diffusion coefficient D(∆, δ), measured perpendicular to major axonal tracts, both on the diffusion time ∆, and on the diffusion gradient pulse width δ. The quantity D(∆, δ) is defined as the lowest-order cumulant term (Kiselev, 2010; Jensen et al., 2005) of the dMRI signal, ln S(∆, δ; g) = −bD(∆, δ) + O(b2) , b = g2δ2(∆ − δ/3) , (1) where g is the applied Larmor frequency gradient, and b is the conventional diffusion weighting (Jones, 2011). The overall D = finDin + fexDex is a weighted average of the appar- ent intra- and extra-axonal diffusivities, with their T2-weighted fractions normalized to fin + fex = 1 (we exclude the contribu- tion of myelin water due to its short T2 ∼ 10 ms (Mackay et al., 1994; Whittall et al., 1997) as compared with our echo time). Remarkably, the functional forms of Din(∆, δ) and Dex(∆, δ) will prove to be sufficiently distinct, enabling us to identify which one dominates. In the limit δ → 0, D(∆, δ)δ→0 → (cid:104)x2(∆)(cid:105)/(2d∆) corre- sponds to the genuine water diffusion coefficient in the d = 2- dimensional plane transverse to the fibers (a weighted average of the genuine compartment diffusivities). Finite-δ measure- ment imposes a low-pass filter (Callaghan, 1991; Burcaw et al., 2015), suppressing the high-frequency dynamics of molecu- lar displacements x(∆); this filter effect is what will techni- cally distinguish Din(∆, δ) and Dex(∆, δ). We will use the ∆-dependence to estimate parameters of both models, and then determine which one predicts the "orthogonal" δ-dependence best. 2. Methods 2.1. Theory We first outline the two models for transverse diffusivity D(∆, δ), paying special attention to their functional forms. In the intra-axonal picture, all ∆- and δ-dependence of the radial diffusivity D ≡ finDin(∆, δ)+fexDex∞ comes from Din, D(∆, δ) (cid:39) D∞ + c δ(∆ − δ/3) , c = 7 48 fin¯r4 D0 (2) based on Neuman's solution (Neuman, 1974) for narrow im- permeable cylinder of radius r (cf. Eq. (A.1) in Appendix A), with the free (axoplasmic) diffusion coefficient D0; for a dis- tribution of axons, the effective ¯r4 ≡ (cid:104)r6(cid:105)/(cid:104)r2(cid:105) (Burcaw et al., 2015). Note that Eq. (2) depends on two independent combi- nations of tissue parameters: c, and the overall bulk diffusion coefficient D∞ = fexDex∞ (in the ∆ → ∞ limit); here Dex∞ is 2 the bulk diffusion coefficient of the extra-axonal water. Typi- cally, δ/3 (cid:28) ∆; in this limit, the ∼ 1/∆ scaling in Eq. (2) is a consequence of a fully restricted geometry. Less obvious, but crucial for our work, is the inverse scaling with the pulse dura- tion, D − D∞ ∼ 1/δ. It can be traced to the intra-axonal dif- fusion attenuation − ln Sin ∝ δ inside a cylinder, being equiv- alent to the effective T ∗ 2 relaxation in the diffusion-narrowing regime (Kiselev and Posse, 1998; Jensen and Chandra, 2000; Sukstanskii and Yablonskiy, 2003, 2004; Novikov and Kiselev, 2008) during the time δ when diffusion gradients are on; the 1/δ scaling follows from factoring out the b ∝ δ2-dependence, cf. Eq. (1). In the extra-axonal picture, attenuation inside axons is ne- glected, i.e. Sin → 1 and Din → 0, and all dependence of D ≡ fexDex on δ and ∆ comes from that of Dex(∆, δ) (Bur- caw et al., 2015; Fieremans et al., 2016): D(∆, δ) (cid:39) D∞ + c(cid:48) · ln(∆/δ) + 3 ∆ − δ/3 c(cid:48) = fexA . (3) 2 , Eq. (3) is again characterized by two combinations of tissue parameters: D∞ and c(cid:48), where D∞ has the same meaning as above, while c(cid:48) is related to the "disorder strength" A charac- terizing the random packing geometry of axons in the extra- axonal space (Fieremans et al., 2016; Burcaw et al., 2015). Here it is crucial that D increases logarithmically with 1/δ, rather than linearly as in Eq. (2). This nontrivial scaling originates from the long-time tail (Novikov et al., 2014; Ernst et al., 1984; Burcaw et al., 2015) of the instantaneous diffusion coefficient 2d ∂t(cid:104)x2(t)(cid:105) (cid:39) Dex∞ + A/t of the extra-axonal wa- Dex ter, restricted by the two-dimensional disordered axonal pack- ing geometry; the gradient pulse width δ provides short-time cutoff for the tail (Fieremans et al., 2016; Burcaw et al., 2015), which can thereby be probed with varying δ. inst(t) = 1 2.2. In vivo MRI Diffusion MRI was performed on five healthy subjects (3 males / 2 females, 25-35 years old), by using a 3T Siemens Prisma scanner (Erlangen, Germany) with a 64-channel head coil. The monopolar pulse-gradient spin-echo (PGSE) diffu- sion tensor imaging (DTI) sequence provided by the vendor (Siemens WIP 511E) was used to perform two different scans for each subject. For each scan, we obtained 3 b = 0 images (no diffusion weighted) and diffusion weighted images (DWI) of b = 0.5 ms/µm2 along 30 diffusion gradient directions, with an isotropic resolution of (2.7 mm)3 and an FOV of (221 mm)2. The scanned brain volume is a slab of 15 slices, aligned parallel to the anterior commissure (AC) to posterior commissure (PC) line. The corpus callosum was in the middle of the slab, such that the entire corpus callosum was scanned (Fieremans et al., 2016). In scan 1, we varied ∆ = [26, 30, 40, 55, 70, 85, 100] ms and fixed δ at 20 ms; in scan 2, we fixed ∆ at 75 ms and varied δ = [4, 5, 6.7, 10, 15, 25, 45] ms. All scans were performed with the same TR/TE = 5000/150 ms. Total acquisition time is ∼ 50 min. 2.3. Image processing Our image processing pipeline includes four steps: denois- ing, Gibbs ringing elimination, eddy-current and motion cor- rection, and diffusion tensor estimation. For denoising, we identified and truncated noise-only prin- ciple components by using the fact that principle component analysis eigenvalues, arising from noise, obey the universal Marchenko-Pastur distribution (Veraart et al., 2016a,c). To elim- inate Gibbs ringing, we re-interpolated each denoised image by sampling the ringing pattern at the zero-crossings of the sinc function (Kellner et al., 2016). Then we used FSL eddy to cor- rect eddy-current distortions and subject motions (Andersson and Sotiropoulos, 2016). Finally, diffusion tensors were evalu- ated via an unconstrained weighted linear least squares (WLLS) method, where the weights were estimated from diffusion ten- sor calculations based on an unweighted LLS method (Veraart et al., 2013). The contribution of imaging gradients to b-value is negligible since it is always less than 10−3 ms/µm2 in our experiments. If the diffusion data have SNR > 2, tensor estimations of WLLS will not be biased by Rician noise (Veraart et al., 2013). To calculate the SNR of b = 0 images, denoised signal was di- vided by the estimated noise level of the noise map, obtained from denoising method mentioned above (Veraart et al., 2016a). In our b = 0 images, mean SNR of the WM was ≈ 18-22. Con- sidering that WM's D(cid:107) ∼ 1.2-1.6 µm2/ms, D ∼ 0.5 µm2/ms and b = 0.5 ms/µm2, SNR in DWIs was still much higher than 2, and thus WLLS gave us unbiased tensor estimations. For each voxel, we calculated eigenvalues of the diffusion tensor estimated via WLLS, sorted in the order λ1 ≥ λ2 ≥ λ3. Axial diffusivity, defined by D(cid:107) ≡ λ1, estimates diffusion parallel to axons. Similarly, radial diffusivity, defined by D ≡ (λ2 + λ3)/2, estimates diffusion transverse to axons. In this way, we obtain maps of axial diffusivity, radial diffusivity, and fractional anisotropy (FA) (Basser et al., 1994). 2.4. Region of Interest (ROI) To automatically delineate WM ROIs, we registered each subject's mean FA map to FSL's standard FA map in MNI 152 space with FMRIB's linear image registration tool (FLIRT) and non-linear registration tool (FNIRT) (Jenkinson and Smith, 2001; Jenkinson et al., 2002; Andersson et al., 2007). The individ- ual mean FA map is acquired by averaging all the FA maps in different ∆ and δ in scans 1 and 2 for each subject. The transformation matrix (FLIRT) and the warp (FNIRT) were re- trieved to inversely transform the WM atlas ROIs from MNI 152 space to the individual subject space. In our study, we used the Johns Hopkins University DTI-based WM atlas (Mori et al., 2005), which was registered to MNI 152 space with FLIRT and FNIRT before use. To suppress the cerebrospinal fluid (CSF) signal contamination due to the long TE, we used an extended CSF mask to exclude WM voxels close to CSF. The CSF mask was segmented from a mean b = 0 image by FMRIB's Auto- mated Segmentation Tool (FAST) (Zhang et al., 2001), and its edge was expanded by one voxel. One subject's WM ROIs are shown in Fig. 1c. In the scanned slab, we focused on the main 3 WM tracts including anterior corona radiata (ACR), superior corona radiata (SCR), posterior corona radiata (PCR), posterior limb of the internal capsule (PLIC), genu, midbody, and sple- nium of the corpus callosum. 2.5. Data Analysis Eigenvalues, axial and radial diffusivities were calculated voxel by voxel and averaged over each ROI. To evaluate the strength of the ∆-dependence described by intra- and extra- axonal models, we assumed that the D in scan 1 is a linear function of 1/(δ(∆ − δ/3)) and (ln(∆/δ) + 3/2)/(∆ − δ/3), suggested by Eq. (2) and Eq. (3), and calculated the two mod- els' Pearson's linear correlation coefficients R and P -values with the null hypothesis of no correlation. If P < 0.05 in an ROI, the null hypothesis is rejected, and the ∆-dependence is non-trivial. In the ROIs with significant ∆-dependence, we fit Eq. (2) and Eq. (3) to the scan 1 data and acquired parameters shown in Table 1. 3. Results In Fig. 1, we show the results for brain scans of five healthy subjects with a monopolar PGSE DTI sequence. The mean val- ues of D were computed within each ROI in brain WM, Fig. 1c, and averaged over five subjects. To explicitly reveal the dependence of D on both ∆ and δ, we performed 2 scans for each subject. In scan 1, we fixed δ = 20 ms, as it is typically done (Fieremans et al., 2016; De San- tis et al., 2016; Barazany et al., 2009; Nilsson et al., 2009; Horsfield et al., 1994; Stanisz et al., 1997; Bar-Shir and Cohen, 2008; Kunz et al., 2013), and varied ∆. Scan 1 embodies a stan- dard t-dependent (t ≈ ∆) dMRI measurement D(t). In scan 2, we fixed ∆ = 75 ms and varied δ instead. This δ-dependence has not been comprehensively studied, and turns out to be quite revealing. Fig. 1a shows that both the intra- and extra-axonal mod- els fit the "standard" scan 1 data well in each ROI. The esti- mated P -value, R2, and fit parameters are shown in Table 1. A naive way to select between the two models would be to use the R2 goodness-of-fit parameter (since both models have the same number of 2 degrees of freedom). However, while R2 is generally closer to 1 for the extra-axonal model, we feel it is not enough to use this noisy metric to unequivocally select Eq. (3). For a physically more informed model selection, we now focus on the functional form of the δ-dependence, by using fit param- eters (D∞ and c, and D∞ and c(cid:48), correspondingly, Table 1), to predict scan 2 data. Fig. 1b shows that the parameter-free predictions of the two models are very different, both quantita- tively and qualitatively; the diffusivity for extra-axonal model, Eq. (3), captures the systematic bend in the curves with respect to 1/δ very well, while Eq. (2) for intra-axonal model increases linearly with 1/δ and clearly deviates from experimental re- sults. We emphasize that the prediction of scan 2 was performed without any adjustable parameters, since tissue properties are found in scan 1, and the δ-dependence is calculated based on P ROI 2.1e-3 ACR 2.5e-4 SCR 6.5e-4 PCR 7.0e-4 PLIC 0.60 Genu 0.24 Midbody Splenium 1.2e-3 R2 0.871 0.945 0.919 0.917 - - D∞ 0.603 0.523 0.592 0.427 - - 43.3 62.3 85.4 80.8 - - 0.896 0.349 106.7 6.43 Intra-axonal model, Eq. (2) (cid:16) fin (cid:17)1/4 c 2¯r ¯η (cid:16) fin (cid:17)1/4 D0 1.73 1.90 2.05 2.03 - - 2.17 P 1.5e-3 1.7e-4 2.9e-4 5.9e-4 0.65 0.31 2.1e-3 √ Extra-axonal model, Eq. (3) l⊥ 1.10 1.30 1.56 1.46 c(cid:48) 0.241 0.338 0.484 0.427 D∞ 0.597 0.515 0.581 0.419 R2 0.887 0.952 0.942 0.922 c fex - - - - - - - - 0.873 0.337 0.560 1.67 D0 5.13 5.62 6.08 6.00 - - Table 1: Estimated parameters from scan 1, based on intra-axonal model, Eq. (2), and extra-axonal model, Eq. (3). Intra-axonal model: Values of 2¯r (fin/D0)1/4 and ¯η (fin/D0)1/4 are lower bounds of, respectively, the (volume-weighted) inner axonal diameter 2¯r (cf. text below Eq. (2)), and of the axonal shrinkage ¯η (Fig. 2 and Eq. (A.2)) since, practically, fin/D0 < 1 ms/µm2. Extra-axonal model: We used empirical estimate (Burcaw et al., 2015) A ≈ 0.2 (l⊥ c )2, to obtain the combination l⊥ c because fex < 1; l⊥ c provides an estimate for the outer axonal diameter. All parameters are in the corresponding units of µm and ms. fex from c(cid:48). This sets a lower bound (Fieremans et al., 2016) on the fiber packing correlation length l⊥ √ c Eq. (2) and Eq. (3). Hence, this prediction provides a parameter- free test of the models involved. Fig. 1 shows that the extra-axonal model demonstrates bet- ter consistency between scans 1 and 2, indicating that the con- tribution of extra-axonal water dominates the signal change. We can also observe this by inspecting model parameter values. Using fit parameters based on the intra-axonal model (see Ta- ble 1, Eq. (2), and Appendix A, Wide pulse limit in the GPA) and typical values of fin ≈ 0.5 and D0 (cid:38) 1.5 µm2/ms (Novikov et al., 2016b), the estimated inner axonal diameter 2¯r ≈ 6.8 − 8.5 µm, much larger than histologically reported values ≈ 1 µm (Aboitiz et al., 1992; Caminiti et al., 2009; Liewald et al., 2014; Tang and Nyengaard, 1997; Tang et al., 1997). Based on fit results of extra-axonal model (see Table 1, Eq. (3), and A ≈ 0.2 (l⊥ c )2 from ref. (Burcaw et al., 2015)), and fex ≈ 0.5, we estimate the axonal packing correlation length c ≈ 1.6 − 2.4 µm. As typical values of the ratio of inner to l⊥ outer diameter (the g-ratio) range within 0.6 − 0.8 in central nervous system (Chomiak and Hu, 2009; Stikov et al., 2015), the outer axonal diameter ∼ 1 µm/(g-ratio) ≈ 1.3 − 1.7 µm, close to estimates of the correlation length in our experiments. Tang and Nyengaard (Tang and Nyengaard, 1997; Tang et al., 1997) uniformly sampled the WM of one human brain hemi- sphere and also reported the outer diameter of myelinated ax- ons of about 1.14 µm on average. The scale of the fiber pack- ing correlation length is biologically plausible, and could be a potential biomarker for the outer diameter, a metric of myeli- nation, which is an important hallmark of neurodegeneration, such as multiple sclerosis (Bando et al., 2015). 4. Discussion By varying both ∆ and δ, and identifying physical origins of these dependencies, our in vivo dMRI measurements dis- tinguish between functional forms of intra- and extra-axonal models, and show the predominance of the extra-axonal time dependence in human brain WM. The extra-axonal model of- fers an estimate of outer axonal diameter via packing correla- tion length, whose changes can be sensitive to demyelination, and possibly axonal loss or other kinds of geometric changes Figure 1: Radial diffusivity D(∆, δ) for WM ROIs averaged over five subjects. (a) With fixed δ = 20 ms, D from scan 1 decreases with ∆. Dashed and solid lines are fits based on Eq. (2) (intra-axonal) and Eq. (3) (extra-axonal), corre- spondingly. (b) With fixed ∆ = 75 ms, D from scan 2 increases as a function of 1/δ. Dashed and solid lines are predictions (not fits) based on parameters obtained from scan 1 (Table 1), using the corresponding models, Eq. (2) and Eq. (3), where now ∆ is fixed and δ varies. (c) WM ROIs, including ACR (red) = anterior corona radiata, SCR (yellow) = superior corona radiata, PCR (green) = posterior corona radiata, PLIC (magenta) = posterior limb of the internal cap- sule, genu (blue), and splenium (cyan) of the corpus callosum. in axonal fiber tracts at the µm level, three orders of magnitude below the achievable resolution of human MRI. In what follows, we will put our work in context of previ- 4 040801200.30.40.50.60.700.10.20.30.30.40.50.60.7 ous measurements using shorter times or thicker axons (in the spinal cord), and employing larger gradients, as well as discuss a possible relation between the disorder strength A character- izing outer axonal diameters, and the measurements of axonal conduction velocity. 4.1. Intra-axonal model: when pulses are not wide Suppose, for a moment, that despite all the above argu- ments, the intra-axonal model is the true one. Then, accord- ing to our results in Table 1, the very large inner axonal radius ¯r ≈ 4µm should lead to an intra-axonal correlation time (time to diffuse across an axon) tc = r2/D0 ≈ 10 ms. Technically, Eq. (2) is applicable only if the wide pulse limit (δ (cid:29) tc) is satisfied (see details in Appendix A). For histologically feasible 2¯r ∼ 1 µm, Eq. (2) applies, since tc < 1 ms, and δ in scan 2 varied from 4-45 ms; that is the reason we used the Neuman's approximation in Eq. (2). However, for the "apparent" tc based on fits of Eq. (2) to scan 1 data, the wide pulse limit is violated. Hence, we will repeat our intra-axonal model analysis using a more general, albeit less analytically transparent equation due to van Gelderen et al. (van Gelderen et al., 1994), applicable to axons of all sizes, and will employ the axonal radius his- togram, Fig. 2, according to histological observations (Caminiti et al., 2009) (cf. Eq. (A.3) and Eq. (A.5) in Appendix A). We will refer to this modified model as the intra-axonal model (van Gelderen). Very large apparent axonal diameters would necessarily im- ply strong brain tissue shrinkage in fixation and paraffin embed- ding (Horowitz et al., 2015b), such that histologically measured axons have to be assumed notably smaller than in vivo. To com- pensate for such hypothetical shrinkage, we introduce a shrink- age factor η > 1, which linearly extends the measured radii his- togram (Fig. 2 and Appendix A), such that mean axonal radius is η times larger than that calculated with histology. We note from the outset, that η cannot exceed 1.5, as argued in refs. (Aboitiz et al., 1992; Houzel et al., 1994), and η ∼ 1.03−1.07 measured in ref. (Tang and Nyengaard, 1997; Tang et al., 1997). Figure 2: Histogram of axonal radii hi = h(ri), based on histological results in corpus callosum of three post-mortem human brains (Caminiti et al., 2009) sampled into 100 bins ri. The shrinkage factor η extends the bins ri → ηri, modeling a correction for the axonal radii due to a uniform tissue shrinkage during fixation, with η = 1 (blue area) corresponding to no shrinkage, i.e. the measured histogram equal to that in vivo. 5 Fig. 3 shows that, in each ROI, the full intra-axonal model (van Gelderen), Eq. (A.3) and Eq. (A.5), can neither fit the scan 1 data nor predict the δ-dependence in scan 2 data if η ≤ 2. (We fixed η to a few values, instead of letting it vary, to achieve fit robustness.) Based on the parameters in Table 2, in most of the ROIs, the values of fin hit the upper bound and the fits are poor (R2 < 0.9) if η ≤ 2. Thus, to fit the data with reason- able parameters and to predict the δ-dependence, the shrinkage due to tissue fixation should exceed two-fold, which contradicts available histological data (Aboitiz et al., 1992; Houzel et al., 1994; Tang and Nyengaard, 1997; Tang et al., 1997). Interestingly, when η > 2, the functional form of the intra- axonal model (van Gelderen) is very similar to that of our extra- axonal model, i.e. the intra-axonal model begins to describe both the varying ∆ and δ data sets equally well. This explains why, in previous studies, which were performed at a few ∆ and δ and which did not focus on the functional form of D(∆, δ), the axonal diameter estimations based on the intra-axonal model alone were much larger than that in histological studies (Alexan- der et al., 2010; Barazany et al., 2009) -- the fitting was "stretch- ing" the axons to match the data. In contrast, the extra-axonal model, Eq. (3), does not stretch the length scales, and provides precise predictions for the δ-dependence, Fig. 1b, and realistic packing correlation length estimates, Table 1. We also note that for the spinal cord, where axons are about factor of 5 thicker than those in the brain, one must use the full van Gelderen's model since tc ∼ 10 ms. In this situation, the balance between the intra- and extra-axonal time-dependencies should be revisited, due to the very strong, ∼ r4 scaling of the intra-axonal signal, so that both effects are now compara- ble. The dMRI measurement can become sensitive to the inner diameters of the spinal cord WM, and reasonable diameter esti- mates can be obtained (Benjamini et al., 2016; Komlosh et al., 2013); however, accounting for the nontrivial time-dependence of the extra-axonal diffusion coefficient still improves such es- timates (Xu et al., 2014). 4.2. Relation to measurements with strong diffusion gradients Applying extremely strong diffusion gradients G ∼ 0.1 − 1 T/m facilitates the estimation of intra-axonal parameters (Barazany et al., 2009; Sepehrband et al., 2016; De Santis et al., 2016; Alexander et al., 2010; Assaf et al., 2008; Huang et al., 2015) because of stronger signal attenuation inside axons, as well as due to exponential suppression of the extra-axonal signal in the radial direction, roughly as ∼ fex e−bDex∞. However, for strong diffusion gradients the intra-axonal model needs corrections, since the Gaussian phase approximation (GPA) for Sin, under which both Neuman's and van Gelderen's solu- tions were obtained, eventually breaks down. Unfortunately, no solutions beyond GPA currently exist for finite pulse width δ. In Appendix A, Beyond GPA, we estimate that GPA breaks down when g (cid:38) g∗ = D0 r3 = 1 r · 1 tc (4) The Larmor frequency gradient g ≡ γG is defined via the pro- ton gyromagnetic ratio γ. For reference, g = 0.0107 (µm · ms)−1 for G = 40 mT/m (typical human scanner). . Figure 3: D from Fig. 1, fit with full intra-axonal model (van Gelderen), Eq. (A.3) and Eq. (A.5), for the four values of the shrinkage factor η = [1, 1.5, 2, 2.4] (dashed lines, top row). Poor fits for η ≤ 2 are due to fin hitting the upper bound. Dashed lines in bottom row are predictions (not fits) for δ-dependence in scan 2 data, based on parameters obtained from scan 1 (Table 2 and Eq. A.5). Solid lines in upper and lower rows are the same as those in Fig. 1, i.e. fits for scan 1 and predictions for scan 2 based on Eq. 3 (extra-axonal model), shown here for reference. ROI η = 1 η = 1.5 Intra-axonal model (van Gelderen), Eq. (A.5) R2 0.145 ACR 0.112 SCR 0.076 PCR 0.071 PLIC Splenium 0.047 * The fitting parameter fin hits the upper bound. D∞ fin 1* 0.611 1* 0.534 1* 0.610 0.440 1* 1* 0.368 R2 0.562 0.463 0.327 0.312 0.211 D∞ fin 1* 0.608 1* 0.531 1* 0.607 0.438 1* 1* 0.366 R2 0.872 0.903 0.733 0.719 0.527 η = 2 D∞ 0.603 0.525 0.600 0.433 0.361 fin 0.901 1* 1* 1* 1* η = 2.4 D∞ 0.602 0.522 0.592 0.427 0.354 R2 0.875 0.950 0.933 0.931 0.785 fin 0.510 0.723 1* 1* 1* Table 2: Estimated parameters from scan 1, based on intra-axonal model (van Gelderen), Eq. (A.5), fixed at four shrinkage factors η. The range of fin is [0, 1]. In most of the ROIs, when η ≤ 2, the fitted fin hits its upper bound, and the fit is poor (R2 < 0.9), which is also shown in the upper rows of Fig. 3. To obtain a better fit and ensure fin < 1, shrinkage factor η needs to exceed 2, which is unrealistic (Aboitiz et al., 1992; Caminiti et al., 2009; Liewald et al., 2014; Houzel et al., 1994; Tang and Nyengaard, 1997; Tang et al., 1997). Estimating Larmor frequency inhomogeneity across an axon by Ω ∼ g∗ · r, the above condition becomes Ω · tc ∼ 1, i.e. the typical precession phase during diffusion across an axon is ∼ 1 (i.e. not small). Note that the critical gradient g∗ is purely determined by tissue properties, independent of sequence tim- ings. For example, if r = 3 µm and D0 = 1.5 µm2/ms, g∗ = 0.0556 (µm · ms)−1 (corresponding to G = 208 mT/m); when 6 the actual g becomes of this order of magnitude (and propor- tionally larger for smaller axons), the higher-order in g correc- tions to GPA become crucial. In our experiments, the gradient strength stays below 77 mT/m, and GPA perfectly applies. Recent studies boosted diffusion gradients up to G (cid:46) 300 mT/m for humans (Huang et al., 2015) and G (cid:46) 1.3 T/m for ex-vivo mice (Sepehrband et al., 2016). 0204060801001200.30.40.50.60.7Fit(Scan1)00.10.20.30.30.40.50.60.7Prediction(Scan2)0204060801001200.30.40.50.60.7Fit(Scan1)00.10.20.30.30.40.50.60.7Prediction(Scan2)0204060801001200.30.40.50.60.7Fit(Scan1)ACRSCRPCRPLICSplenium00.10.20.30.30.40.50.60.7Prediction(Scan2)0204060801001200.30.40.50.60.7Fit(Scan1)ACRSCRPCRPLICSplenium00.10.20.30.30.40.50.60.7Prediction(Scan2) When the signal contribution of large axons is not negligible, beyond-GPA corrections are needed due to the tail of axonal histogram extending to large axons, since for them, the criti- cal g∗ decreases as 1/r. The negative beyond-GPA correction to ln Sin, Eq. (A.7), may therefore explain the residual overes- timation of axonal diameters in the study (Sepehrband et al., 2016) with ultra-strong gradients -- basically, this correction tells that Sin experiences extra attenuation due to the O(g4) contribution, neglected in standard axonal diameter mapping frameworks. Similarly, higher-order corrections in the powers of diffu- sion weighting b ∝ g2, Eq. (1), should be considered for the extra-axonal signal. The extra-axonal signal Sex up to O(b2) can be obtained from the recent narrow-pulse result [Appendix E of ref. (Burcaw et al., 2015)], by substituting tc → δ as the logarithmic cutoff: ln Sex (cid:39) −b Dex(∆, δ) + Kex 6 (bDex∞)2 , (5) where Kex(∆, δ) is the apparent extra-axonal kurtosis, , ∆ (cid:29) δ (cid:29) tc . · ln(∆/δ) Kex(∆, δ) 6 (cid:39) A Dex∞ ∆ Here, the genuine kurtosis Kex(t) has the ln(t/tc)/t tail (Bur- caw et al., 2015), and we used the low-pass filter analogy in the wide pulse limit δ (cid:29) tc, to re-define the long-time tail cut-off, tc → δ. Eq. (5) tells that the O(b2) kurtosis term becomes of the or- der of the nontrivial, time-dependent O(b) term, when bDex∞ (cid:38) 1; this condition practically coincides with the breakdown of the O(b), DTI representation ln S ≈ −b · fexDex∞, for the total signal S (cid:39) fin +fex e−bDex∞. In other words, at the same b when the curvature of the observed ln S versus b becomes notable, the extra-axonal Kex term in Eq. (5) must be included in the anal- ysis if one wants to estimate A and fin, fex (and, possibly the inner radii) separately, by going to high b; one cannot use the approximation SKex≡0 (cid:39) finSin + fex e−bDex(∆,δ) beyond its O(b) term. De Santis et al. (De Santis et al., 2016) used the SKex≡0 approximation to modify AxCaliber estimation of in- ner diameters from human brain data in the corpus callosum acquired with stimulated echo dMRI with b ≤ 4 ms/µm2. In- cluding the Dex(∆, δ) term in Eq. (5) resulted in about 5-fold smaller inner diameter estimates in comparison to just using Dex∞, effectively demonstrating the importance of non-Gaussian (time-dependent) extra-axonal space contribution to the total signal, consistent with ref. (Fieremans et al., 2016). However, δ was fixed to a single value while ∆ varied, even though Sin mostly depends on δ, and ∆-dependence drops out in the Neu- man's limit, cf. Appendix A. Omission of the equally important Kex contribution (as well as, possibly, higher-order cumulant terms) has introduced an unknown bias into parameter estima- tion. Here, we limited our analysis to b ≤ 0.5 ms/µm2 to stay in the linear, DTI regime of Eq. (1). We therefore cannot esti- mate A and compartment fractions separately; such estimation would require a systematic measurement of both ∆ and δ de- pendencies at higher b, and including higher-order cumulants 7 into the model for Sex(∆, δ; b). This is beyond the scope of the present work. We also attempted to fit to scan 1 data a hybrid model D = finDin(∆, δ) + fexDex(∆, δ), including fi- nite axonal radius histogram in the van Gelderen's framework of Din(∆, δ); fitting results were unstable, and corresponding parameters were highly dependent on their initial values, sig- nifying a "shallow direction" in the fitting landscape. Such spurious parameter correlation should be expected from simi- lar functional forms of the extra-axonal model and of the intra- axonal (van Gelderen) model for large inner radii, cf. Fig. 3 for large η. To simplify models and interpretations, we ignored the fiber orientation dispersion, which is generally non-negligible in the brain WM (Zhang et al., 2011; Alexander et al., 2010; Ver- aart et al., 2016b). In the future, it may be possible to consis- tently factor out this dispersion by generalizing the rotationally- invariant parameter estimation (Novikov et al., 2016b; Reisert et al., 2016) onto time-dependent diffusion propagators. 4.3. Correlation of dMRI with axonal conduction velocity Generally, thicker axons have higher axonal conduction ve- locity (ACV). Within the neuroscience community, it is still an open question whether it is inner or outer axonal diame- ter, or some combination of both, that determine ACV most definitively. Hursh (Hursh, 1939) observed that, in the periph- eral nerve of cats and kittens, the ACV was linearly correlated with the outer axonal diameter. Rushton, and Waxman & Ben- nett (Waxman and Bennett, 1972; Rushton, 1951) reanalyzed Hursh's data, and all concluded that ACV is proportional to the outer diameter. However, Sanders and Whitteridge's results in rabbit's peroneal nerve showed that the myelin sheath thick- ness, i.e. the difference between outer and inner radii, had the highest correlation with ACV, rather than inner and outer diam- eters separately (Sanders and Whitteridge, 1946). Arbuthnott et al. (Arbuthnott et al., 1980) studied the peripheral nerve of cat and suggested that conduction velocity is proportional to inner diameter; however, they did not measure the conduction velocity in this study, and the conclusion was made based on their theoretical discussion. To estimate the ACV in the human brain, Aboitiz et al. assumed that inner diameter has a linear re- lationship with ACV; the proportionality constant is 8.7 mm/ms per µm of inner diameter, which is calculated in the peripheral nervous system (Aboitiz et al., 1992; Ruch and Patton, 1982). The advent of in vivo dMRI has offered an exciting proposi- tion to map axonal diameters, and to study in vivo the decades- old relation between axonal sizes and ACV. In 2014, based on the AxCaliber interpretation of dMRI, Horowitz et al. (Horowitz et al., 2015a) estimated apparent inner axonal diameters in the in vivo human brain, and displayed their correlation with ACV measured with electroencephalography; the estimated propor- tionality constant was close to the value used by Aboitiz et al. (Aboitiz et al., 1992; Ruch and Patton, 1982) The finding was subsequently criticized by Innocenti, Caminiti and Aboitiz (Innocenti et al., 2015) since the estimated inner diameter was much larger than histological observations, and the measured interhemispheric transfer time was much shorter than the value in previous literature. This debate presents an interesting sci- entific question: Can one rationalize fairly strong apparent cor- relations between dMRI and ACV observed by Horowitz et al. (Horowitz et al., 2015a) with the inconsistencies of inner diam- eter estimation methodology? The relevance of the nontrivial dMRI signal from the extra- axonal space leads us to posit that the correlation uncovered by Horowitz et al. (Horowitz et al., 2015a) is, to the leading or- der O(b), between the strength of time dependence (c or c(cid:48) in Eq. (2) or Eq. (3)), and ACV. Interpreting the strength of time dependence as inner diameter or extra-axonal packing correla- tion length depends on the model selection. Our present model selection results suggest re-interpreting dMRI axonal diameter mapping in terms of the dominant extra-axonal contribution, defined in terms of the "disorder strength" A, and the related axonal packing correlation length l⊥ c(cid:48) estimating outer diameters. Selecting the extra-axonal model based on our current data is then consistent with the above mentioned cor- relations(Hursh, 1939; Waxman and Bennett, 1972; Rushton, 1951; Sanders and Whitteridge, 1946) between, predominantly, the outer axonal diameters and ACV. c ∼ √ A ∼ √ 5. Conclusions We considered the functional form of D(∆, δ) for two plau- sible biophysical models with mutually exclusive physical as- sumptions. We experimentally showed in the in vivo human brain, that the extra-axonal model provides a far better agree- ment with the measurement, both in terms of the quality of its parameter-free prediction of the measurement with varying δ, and in terms of the qualitative ln(1/δ), rather than 1/δ, func- tional form. Varying δ has revealed a nontrivial low-pass filter effect of the gradient duration on the genuine molecular diffu- sion coefficient D(t). Extra-axonal model provides reasonable values of the pack- ing correlation length, which is compatible to the scale of outer axonal diameter. In contrast, intra-axonal model alone overes- timates the inner axonal diameters by at least twofold as com- pared with histology, which cannot be explained by any reason- able degree of the tissue shrinkage in fixation. The sensitivity of time-dependent diffusion to packing ge- ometry of the extra-axonal space may serve as a marker for de- myelination or axonal loss in neurodegenerative diseases. Our results are also consistent with the correlations between outer axonal diameter and axonal conduction velocity. Acknowlegements We thank Thorsten Feiweier for developing advanced diffu- sion WIP sequence and Jelle Veraart for assistance in process- ing. Research was supported by the National Institute of Neu- rological Disorders and Stroke of the NIH under award number R01NS088040. 8 Appendix A. Intra-axonal model Here we obtain qualitative estimates for signal attenuation within an impermeable cylinder in the GPA, outline exact rela- tions for Din(∆, δ) in different limits, and estimate when GPA breaks down. Mapping onto transverse relaxation Fundamentally, dMRI is a measurement of transverse NMR relaxation in the applied diffusion gradient. Each spin, follow- ing its Brownian path x(τ ), contributes the precession phase 0 Ω(cid:0)x(τ ), τ(cid:1) dτ, where Ω(x, τ ) is the local e−iφ(t), φ(t) = (cid:82) t Larmor frequency offset (relative to γB0), that also depends on time τ explicitly due to the time-varying applied gradient. The dMRI signal S = (cid:104)e−iφ(cid:105) ≡ p(λ)λ=1, given by the average over all spins in a voxel, is, effectively, the Fourier transform p(λ) = (cid:104)e−iλφ(cid:105) of the probability density function P(φ) of all possible precession phases φ(t), where (cid:104). . .(cid:105) is the average with respect to P(φ). Wide-pulse limit in the GPA as φ ∼ (cid:80)N Generally, the form of P(φ) is quite complicated, and is mediated by the diffusion (Kiselev and Posse, 1998; Jensen and Chandra, 2000; Sukstanskii and Yablonskiy, 2003, 2004; Novikov and Kiselev, 2008). Fortunately, in the wide-pulse limit δ (cid:29) tc, the problem of finding its Fourier transform p(λ) simplifies, as the problem maps onto that of transverse relax- ation in the diffusion-narrowing regime (equivalent to the GPA). In this limit, the time tc to diffuse across an axon of radius r provides the correlation time, beyond which the contribution to the precession phase φ for each spin gets randomized. It is then natural to split each Brownian path x(τ ) into N = t/tc (cid:29) 1 steps of duration tc, such that the total phase can be estimated n=1 φn, where each φn ∼ Ω · tc can be treated as an independent random variable with zero mean and variance n(cid:105) ∼ (Ω · tc)2; here Ω ∼ g · r is a typical value of the (cid:104)φ2 Larmor frequency inhomogeneity across an axon imposed by the applied gradient g. When the number N of independent "steps" becomes large, the Central limit theorem (CLT) tells that the characteristic function p(λ) (cid:39) e−iλ(cid:104)φ(cid:105)−λ2(cid:104)φ2(cid:105)c/2 ap- proaches that of the Gaussian distribution, with the higher-order cumulants being less relevant. Moreover, according to the CLT, the mean values and variances from the independent steps add up, i.e. (cid:104)φ(cid:105) ≡ 0, and (cid:104)φ2(cid:105)c ≡ (cid:104)φ2(cid:105) − (cid:104)φ(cid:105)2 ∼ N(cid:104)φ2 n(cid:105) ∼ 2 ∼ Ω2 tc, Ω2 tc · t, such that Sin ∼ e−R∗ cf. refs. (Kiselev and Posse, 1998; Jensen and Chandra, 2000; Sukstanskii and Yablonskiy, 2003, 2004; Novikov and Kiselev, 2008). In our case, it is the total pulse duration t = 2δ that matters; note that the inter-pulse duration ∆ ≥ δ does not enter these considerations, as long as δ (cid:29) tc, since Ω(x, τ ) ≡ 0 and no transverse relaxation occurs during the time when the gradi- ent is off. Hence, the O(g2) attenuation inside an axon scales 2(cid:104)φ2(cid:105)c ∼ (g2r4/D0) · δ, which indeed agrees as − ln Sin (cid:39) 1 with the 1974 exact calculation of Neuman (Neuman, 1974) 2·t, with effective R∗ − ln Sin = 7 48 · g2r4 D0 · δ + O(g4) , (A.1) 9 where the coefficient 7/48 is specific to the assumed perfectly circular cylinder cross-section. Factoring out b in Eq. (A.1), cf. Eq. (1), leads to the intra- axonal contribution in Eq. (2). The corresponding Din is about 2 − 5 × 10−4 µm2/ms for r ∼ 1 µm, D0 = 1.5 µm2/ms, and δ = 20 ms, being much smaller than the measured diffusivity change in our experiment; to account for the observed diffusiv- ity variation over diffusion times, apparent radii ¯r need to be much larger, cf. Table 1. The estimated shrinkage factor for apparent radii ¯r in the Neuman's regime is ¯η ≡ ¯r , where the denominator is the apparent radius(cid:0)(cid:104)r6(cid:105)/(cid:104)r2(cid:105)(cid:1) 1 4 cal- culated via the histology histogram (Caminiti et al., 2009), the blue area in Fig. 2. 1.48 µm (A.2) General solution in the GPA When Neuman's assumption δ (cid:29) tc is not satisfied, one needs to use the general O(g2) solution for signal attenuation inside a cylinder of radius r by van Gelderen et al. (van Gelderen et al., 1994): (cid:20) − ln SvG in (∆, δ; r) = +2e−α2 mδ/tc + 2e−α2 2g2r4 D0 tc m − 1) α6 m(α2 m∆/tc − e−α2 m(∆−δ)/tc − e−α2 m=1 · 2α2 m − 2 δ tc m(∆+δ)/tc (cid:105) (A.3) where αm is the mth root of dJ1(α)/dα = 0, and J1(α) is the Bessel function of the first kind; note that tc = tc(r) = r2/D0. In the δ (cid:29) tc limit, the ∆-dependence drops out, and Eq. (A.3) approaches Eq. (A.1). In the opposite, narrow-pulse limit δ (cid:28) tc, ln Sin(t, δ; r)δ=0 = −bD(t), with D(t) = r2/(4t). In our analysis, we incorporate the axonal radius histogram from the corpus callosum of three post-mortem human brains, by Caminiti et al. (Caminiti et al., 2009), Fig. 2, and allow for the uniform axonal stretching, ri → ηri, such that the overall intra-axonal signal for a given shrinkage factor η is the volume- averaged Eq. (A.3) Sin(∆, δ; η) = fi SvG in (∆, δ; ηri) (A.4) j hjr2 in with the normalized weights fi = hir2 j given in terms (∆, δ; η) is of the histogram bin values hi. The effective DvGel obtained by factoring out the b-value from ln Sin, cf. Eq. (1). The average radius is (cid:104)r(cid:105) ∼ 0.67 µm × η according to the weights hi. The value of g2 is estimated by the b-value de- fined in Eq. (1). The intra-axonal model based on van Gelderen et al. 's solution then yields DvGel(∆, δ; η) = D∞ + fin DvGel in (∆, δ; η) . (A.5) This model includes four parameters: D∞, fin, η, and D0. To stabilize our fitting, we fixed D0 by the value of the axial diffu- sivity D(cid:107) from the diffusion tensor, and fixed η at a few values ∞(cid:88) (cid:88) i i /(cid:80) [1, 1.5, 2, 2.4]. After that, we only have two fitted parameters, D∞ and fin, estimated from scan 1 data, in Table 2. Using these parameters, we predicted the δ-dependence in scan 2 re- sults based on Eq. (A.5) without tunable parameters, shown in Fig. 3. Beyond GPA Unfortunately, there are no exact results for the O(g4) terms and beyond in Eq. (A.1). Let us estimate this next-order term using similar qualitative considerations as above, and establish where the GPA breaks down. For that, we need to estimate the 4th-order cumulant (cid:104)φ4(cid:105)c ≡ (cid:104)φ4(cid:105) − 3(cid:104)φ2(cid:105)2 of the precession phase in the cumulant expansion (Kiselev, 2010) of p(λ) taken at λ ≡ 1: 1 4! (cid:104)φ4(cid:105)c − . . . . ln S = − 1 2! (cid:104)φ2(cid:105)c + (A.6) By definition of the kurtosis K of the phase distribution P(φ), (cid:104)φ4(cid:105)c = K · (cid:104)φ2(cid:105)2 In the large-N limit, kurtosis scales as c. K ∼ −1/N ∼ −tc/δ and is negative as a result of the confined intra-axonal geometry (see the derivation in Supplementary In- formation, Section I). As a result, we obtain · δ . (A.7) GPA breaks down when (cid:104)φ4(cid:105)c ∼ (cid:104)φ2(cid:105)c in Eq. (A.6), equivalent to (cid:104)φ2(cid:105)c ∼ δ/tc, from which the breakdown condition, Eq. (4) in the main text, follows. For such strong gradients, all terms in cumulant expansion are of the same order, which requires development of non-perturbative approaches. (cid:104)φ4(cid:105)c ∼ − g4r10 D3 0 References Aboitiz, F., Scheibel, A. B., Fisher, R. S., Zaidel, E., 1992. Fiber composition of the human corpus callosum. Brain Res 598 (1-2), 143 -- 153. URL http://www.ncbi.nlm.nih.gov/pubmed/1486477 Ackerman, J. J. H., Neil, J. J., Aug 2010. The use of MR-detectable reporter molecules and ions to evaluate diffusion in normal and ischemic brain. NMR Biomed 23 (7), 725 -- 33. Alexander, D. C., Hubbard, P. L., Hall, M. G., Moore, E. A., Ptito, M., Parker, G. J., Dyrby, T. B., 2010. Orientationally invariant indices of axon diameter and density from diffusion MRI. Neuroimage 52 (4), 1374 -- 1389. URL http://www.ncbi.nlm.nih.gov/pubmed/20580932 Andersson, J. L., Jenkinson, M., Smith, S., 2007. Non-linear registration, aka Spatial normalisation FMRIB technical report TR07JA2. Report, FMRIB Analysis Group of the University of Oxford. Andersson, J. L., Sotiropoulos, S. N., 2016. An integrated approach to correc- tion for off-resonance effects and subject movement in diffusion MR imag- ing. Neuroimage 125, 1063 -- 1078. URL http://www.ncbi.nlm.nih.gov/pubmed/26481672 Arbuthnott, E. R., Boyd, I. A., Kalu, K. U., 1980. Ultrastructural dimensions of myelinated peripheral nerve fibres in the cat and their relation to conduction velocity. J Physiol 308, 125 -- 157. URL http://www.ncbi.nlm.nih.gov/pubmed/7230012 Assaf, Y., Basser, P. J., 2005. Composite hindered and restricted model of diffusion (CHARMED) MR imaging of the human brain. NeuroImage 27 (1), 48 -- 58. URL pii/S1053811905002259 http://www.sciencedirect.com/science/article/ Assaf, Y., Blumenfeld-Katzir, T., Yovel, Y., Basser, P. J., 2008. AxCaliber: a method for measuring axon diameter distribution from diffusion MRI. Magn Reson Med 59 (6), 1347 -- 1354. URL http://www.ncbi.nlm.nih.gov/pubmed/18506799 10 Bando, Y., Nomura, T., Bochimoto, H., Murakami, K., Tanaka, T., Watanabe, T., Yoshida, S., 2015. Abnormal morphology of myelin and axon pathology in murine models of multiple sclerosis. Neurochemistry International 81, 16 -- 27. URL pii/S0197018615000066 http://www.sciencedirect.com/science/article/ Bar-Shir, A., Cohen, Y., 2008. High b-value q-space diffusion mrs of nerves: structural information and comparison with histological evidence. NMR in Biomedicine 21 (2), 165 -- 174. URL http://dx.doi.org/10.1002/nbm.1175 Barazany, D., Basser, P., Assaf, Y., 2009. In vivo measurement of axon diameter distribution in the corpus callosum of rat brain. Brain 132, 1210 -- 1220. Basser, P. J., Mattiello, J., LeBihan, D., 1994. MR diffusion tensor spec- troscopy and imaging. Biophysical Journal 66 (1), 259 -- 267. URL 80775-1 http://dx.doi.org/10.1016/S0006-3495(94) Benjamini, D., Komlosh, M. E., Holtzclaw, L. A., Nevo, U., Basser, P. J., 2016. White matter microstructure from nonparametric axon diameter distribution mapping. NeuroImage 135, 333 -- 344. URL pii/S1053811916300921 http://www.sciencedirect.com/science/article/ Burcaw, L. M., Fieremans, E., Novikov, D. S., 2015. Mesoscopic structure of neuronal tracts from time-dependent diffusion. Neuroimage 114, 18 -- 37. URL http://www.ncbi.nlm.nih.gov/pubmed/25837598 Callaghan, P. T., 1991. Principles of Nuclear Magnetic Resonance Microscopy. Clarendon, Oxford. Caminiti, R., Ghaziri, H., Galuske, R., Hof, P. R., Innocenti, G. M., 2009. Evo- lution amplified processing with temporally dispersed slow neuronal con- nectivity in primates. Proc Natl Acad Sci USA 106 (46), 19551 -- 19556. URL http://www.ncbi.nlm.nih.gov/pubmed/19875694 Chomiak, T., Hu, B., 2009. What is the optimal value of the g-ratio for myeli- nated fibers in the rat CNS? a theoretical approach. PLoS One 4 (11), e7754. URL http://www.ncbi.nlm.nih.gov/pubmed/19915661 De Santis, S., Jones, D. K., Roebroeck, A., 2016. Including diffusion time de- pendence in the extra-axonal space improves in vivo estimates of axonal diameter and density in human white matter. Neuroimage 130, 91 -- 103. URL http://www.ncbi.nlm.nih.gov/pubmed/26826514 Ernst, M. H., Machta, J., Dorfman, J. R., van Beijeren, H., 1984. Long-time tails in stationary random media. 1. Theory. Journal of Statistical Physics 34 (3-4), 477 -- 495. Fieremans, E., Burcaw, L. M., Lee, H. H., Lemberskiy, G., Veraart, J., Novikov, D. S., 2016. In vivo observation and biophysical interpretation of time- dependent diffusion in human white matter. Neuroimage 129, 414 -- 427. URL http://www.ncbi.nlm.nih.gov/pubmed/26804782 Horowitz, A., Barazany, D., Tavor, I., Bernstein, M., Yovel, G., Assaf, Y., 2015a. In vivo correlation between axon diameter and conduction velocity in the human brain. Brain Struct Funct 220 (3), 1777 -- 1788. URL http://www.ncbi.nlm.nih.gov/pubmed/25139624 Horowitz, A., Barazany, D., Tavor, I., Yovel, G., Assaf, Y., 2015b. Response to the comments on the paper by horowitz et al.(2014). Brain Structure and Function 220 (3), 1791. Horsfield, M. A., Barker, G. J., McDonald, W. I., 1994. Self-diffusion in cns tissue by volume-selective proton nmr. Magnetic Resonance in Medicine 31 (6), 637 -- 644. URL http://dx.doi.org/10.1002/mrm.1910310609 Houzel, J. C., Milleret, C., Innocenti, G., 1994. Morphology of callosal axons interconnecting areas 17 and 18 of the cat. Eur J Neurosci 6 (6), 898 -- 917. URL http://www.ncbi.nlm.nih.gov/pubmed/7952278 Huang, S. Y., Nummenmaa, A., Witzel, T., Duval, T., Cohen-Adad, J., Wald, L. L., McNab, J. A., 2015. The impact of gradient strength on in vivo diffu- sion mri estimates of axon diameter. Neuroimage 106, 464 -- 472. URL http://www.ncbi.nlm.nih.gov/pubmed/25498429 Hursh, J. B., 1939. Conduction velocity and diameter of nerve fibers. American Journal of Physiology 127 (1), 131 -- 139. URL <GotoISI>://WOS:000202435900013 Innocenti, G. M., Caminiti, R., Aboitiz, F., 2015. Comments on the paper by Horowitz et al. (2014). Brain Struct Funct 220 (3), 1789 -- 1790. URL http://www.ncbi.nlm.nih.gov/pubmed/25579065 Jenkinson, M., Bannister, P., Brady, M., Smith, S., 2002. Improved optimiza- tion for the robust and accurate linear registration and motion correction of brain images. Neuroimage 17 (2), 825 -- 841. URL http://www.ncbi.nlm.nih.gov/pubmed/12377157 Jenkinson, M., Smith, S., 2001. A global optimisation method for robust affine registration of brain images. Med Image Anal 5 (2), 143 -- 156. URL http://www.ncbi.nlm.nih.gov/pubmed/11516708 Jensen, J. H., Chandra, R., Jul 2000. Nmr relaxation in tissues with weak mag- netic inhomogeneities. Magn Reson Med 44 (1), 144 -- 56. Jensen, J. H., Helpern, J. A., Ramani, A., Lu, H., Kaczynski, K., 2005. Diffu- sional kurtosis imaging: The quantification of non-gaussian water diffusion by means of magnetic resonance imaging. Magnetic Resonance in Medicine 53 (6), 1432 -- 1440. URL http://dx.doi.org/10.1002/mrm.20508 Jones, D. K., 2011. Diffusion MRI: Theory, Methods, and Applications. Oxford University Press, New York. Kellner, E., Dhital, B., Kiselev, V. G., Reisert, M., 2016. Gibbs-ringing artifact removal based on local subvoxel-shifts. Magn Reson Med 76 (5), 1574 -- 1581. URL http://www.ncbi.nlm.nih.gov/pubmed/26745823 Kiselev, V. G., 2010. The cumulant expansion: an overarching mathematicl framework for understanding diffusion NMR. In: Jones, D. (Ed.), Diffusion MRI: theory, methods, and applications. Oxford University Press, Ch. 10, pp. 152 -- 168. Kiselev, V. G., 2017. Fundamentals of diffusion MRI physics. NMR in Biomedicine DOI:10.1002/nbm.3602. URL http://dx.doi.org/10.1002/nbm.3602 Kiselev, V. G., Posse, S., 1998. Analytical theory of susceptibility induced nmr signal dephasing in a cerebrovascular network. Physical Review Let- ters 81 (25), 5696. Komlosh, M., A-zarslan, E., Lizak, M., Horkayne-Szakaly, I., Freidlin, R., Horkay, F., Basser, P., 2013. Mapping average axon diameters in porcine spinal cord white matter and rat corpus callosum using d-pfg {MRI}. NeuroImage 78, 210 -- 216. URL pii/S1053811913003273 http://www.sciencedirect.com/science/article/ Kunz, N., Sizonenko, S. V., Huppi, P. S., Gruetter, R., van de Looij, Y., 2013. In- vestigation of field and diffusion time dependence of the diffusion-weighted signal at ultrahigh magnetic fields. NMR in Biomedicine 26 (10), 1251 -- 1257. URL http://dx.doi.org/10.1002/nbm.2945 Liewald, D., Miller, R., Logothetis, N., Wagner, H. J., Schuz, A., 2014. Distri- bution of axon diameters in cortical white matter: an electron-microscopic study on three human brains and a macaque. Biol Cybern 108 (5), 541 -- 557. URL http://www.ncbi.nlm.nih.gov/pubmed/25142940 Mackay, A., Whittall, K., Adler, J., Li, D., Paty, D., Graeb, D., 1994. In vivo visualization of myelin water in brain by magnetic resonance. Magnetic Res- onance in Medicine 31 (6), 673 -- 677. URL http://dx.doi.org/10.1002/mrm.1910310614 Mitra, P. P., Sen, P. N., Schwartz, L. M., Le Doussal, P., June 1992. Diffusion propagator as a probe of the structure of porous media. Physical Review Letters 68 (24), 3555 -- 3558. Mori, S., Wakana, S., Van Zijl, P. C., Nagae-Poetscher, L., 2005. MRI atlas of human white matter. Elsevier, Amsterdam, The Netherlands. Neuman, C. H., 1974. Spin-echo of spins diffusing in a bounded medium. Jour- nal of Chemical Physics 60 (11), 4508 -- 4511. URL <GotoISI>://WOS:A1974T286300056 Nilsson, M., L Att, J., Nordh, E., Wirestam, R., St Ahlberg, F., Brockstedt, S., 2009. On the effects of a varied diffusion time in vivo: is the diffusion in white matter restricted? Magnetic Resonance Imaging 27 (2), 176 -- 187. URL pii/S0730725X08002014 http://www.sciencedirect.com/science/article/ Novikov, D. S., Jensen, J. H., Helpern, J. A., Fieremans, E., 2014. Revealing mesoscopic structural universality with diffusion. Proc Natl Acad Sci USA 111 (14), 5088 -- 5093. URL http://www.ncbi.nlm.nih.gov/pubmed/24706873 Novikov, D. S., Jespersen, S. N., Kiselev, V. G., Fieremans, E., 2016a. Quanti- fying brain microstructure with diffusion MRI: Theory and parameter esti- mation. preprint arXiv:1612.02059. URL http://arxiv.org/abs/1612.02059 Novikov, D. S., Kiselev, V. G., Nov 2008. Transverse NMR relaxation in mag- netically heterogeneous media. J Magn Reson 195 (1), 33 -- 9. Novikov, D. S., Veraart, J., Jelescu, I. O., Fieremans, E., 2016b. Mapping ori- entational and microstructural metrics of neuronal integrity with in vivo dif- fusion MRI. preprint arXiv:1609.09144 https://arxiv.org/abs/1609.09144. Reisert, M., Kellner, E., Dhital, B., Hennig, J., Kiselev, V. G., 2016. Disentan- gling Micro from Mesostructure by diffusion MRI: A Bayesian Approach. NeuroImage. Reynaud, O., Winters, K. V., Hoang, D. M., Wadghiri, Y. Z., Novikov, D. S., Kim, S. G., Jul 2016. Surface-to-volume ratio mapping of tumor microstruc- ture using oscillating gradient diffusion weighted imaging. Magn Reson Med 76 (1), 237 -- 47. Ruch, T., Patton, H., 1982. Physiology and Biophysics. Vol. 4. Saunders, Philadelphia. Rushton, W. A., 1951. A theory of the effects of fibre size in medullated nerve. J Physiol 115 (1), 101 -- 122. URL http://www.ncbi.nlm.nih.gov/pubmed/14889433 Sanders, F. K., Whitteridge, D., 1946. Conduction velocity and myelin thick- ness in regenerating nerve fibres. J Physiol 105, 152 -- 174. URL http://www.ncbi.nlm.nih.gov/pubmed/20999939 Sepehrband, F., Alexander, D. C., Kurniawan, N. D., Reutens, D. C., Yang, Z., 2016. Towards higher sensitivity and stability of axon diameter estimation with diffusion-weighted MRI. NMR Biomed 29 (3), 293 -- 308. URL http://www.ncbi.nlm.nih.gov/pubmed/26748471 Stanisz, G. J., Wright, G. A., Henkelman, R. M., Szafer, A., 1997. An analytical model of restricted diffusion in bovine optic nerve. Magnetic Resonance in Medicine 37 (1), 103 -- 111. URL http://dx.doi.org/10.1002/mrm.1910370115 Stikov, N., Campbell, J. S., Stroh, T., Lavel´ee, M., Frey, S., Novek, J., Nuara, S., Ho, M.-K., Bedell, B. J., Dougherty, R. F., Leppert, I. R., Boudreau, M., Narayanan, S., Duval, T., Cohen-Adad, J., Picard, P.-A., Gasecka, A., Cot´e, D., Pike, G. B., 2015. In vivo histology of the myelin g-ratio with magnetic resonance imaging. NeuroImage 118, 397 -- 405. URL pii/S1053811915004036 http://www.sciencedirect.com/science/article/ Sukstanskii, A. L., Yablonskiy, D. A., Aug 2003. Gaussian approximation in the theory of mr signal formation in the presence of structure-specific magnetic field inhomogeneities. J Magn Reson 163 (2), 236 -- 47. Sukstanskii, A. L., Yablonskiy, D. A., Mar 2004. Gaussian approximation in the theory of mr signal formation in the presence of structure-specific mag- netic field inhomogeneities. effects of impermeable susceptibility inclusions. J Magn Reson 167 (1), 56 -- 67. Tang, Y., Nyengaard, J., Pakkenberg, B., Gundersen, H., 1997. Age-induced white matter changes in the human brain: A stereological investigation. Neurobiology of Aging 18 (6), 609 -- 615. URL pii/S0197458097001553 http://www.sciencedirect.com/science/article/ Tang, Y., Nyengaard, J. R., 1997. A stereological method for estimating the total length and size of myelin fibers in human brain white matter. J Neurosci Methods 73 (2), 193 -- 200. URL http://www.ncbi.nlm.nih.gov/pubmed/9196291 Tanner, J., 1979. Self diffusion of water in frog muscle. Biophysical journal 28 (1), 107. van Gelderen, P., DesPres, D., van Zijl, P. C., Moonen, C. T., 1994. Evaluation of restricted diffusion in cylinders. Phosphocreatine in rabbit leg muscle. J Magn Reson B 103 (3), 255 -- 260. URL http://www.ncbi.nlm.nih.gov/pubmed/8019777 Veraart, J., Fieremans, E., Novikov, D. S., 2016a. Diffusion MRI noise mapping using random matrix theory. Magn Reson Med 76 (5), 1582 -- 1593. URL http://www.ncbi.nlm.nih.gov/pubmed/26599599 Veraart, J., Fieremans, E., Novikov, D. S., 2016b. Universal power-law scaling of water diffusion in human brain defines what we see with MRI. preprint arXiv:1609.09145 https://arxiv.org/abs/1609.09145. Veraart, J., Novikov, D. S., Christiaens, D., Ades-aron, B., Sijbers, J., Fiere- mans, E., 2016c. Denoising of diffusion {MRI} using random matrix theory. NeuroImage 142, 394 -- 406. URL pii/S1053811916303949 http://www.sciencedirect.com/science/article/ Veraart, J., Sijbers, J., Sunaert, S., Leemans, A., Jeurissen, B., 2013. Weighted linear least squares estimation of diffusion MRI parameters: strengths, lim- itations, and pitfalls. Neuroimage 81, 335 -- 346. URL http://www.ncbi.nlm.nih.gov/pubmed/23684865 Waxman, S. G., Bennett, M. V., 1972. Relative conduction velocities of small myelinated and non-myelinated fibres in the central nervous system. Nat New Biol 238 (85), 217 -- 219. 11 URL http://www.ncbi.nlm.nih.gov/pubmed/4506206 Whittall, K. P., MacKay, A. L., Graeb, D. A., Nugent, R. A., Li, D. K., Paty, D. W., 1997. In vivo measurement of T2 distributions and water contents in normal human brain. Magn Reson Med 37 (1), 34 -- 43. URL http://www.ncbi.nlm.nih.gov/pubmed/8978630 Xu, J., Li, H., Harkins, K. D., Jiang, X., Xie, J., Kang, H., Does, M. D., Gore, J. C., 2014. Mapping mean axon diameter and axonal volume fraction by MRI using temporal diffusion spectroscopy. NeuroImage 103, 10 -- 19. Yablonskiy, D. A., Sukstanskii, A. L., 2010. Theoretical models of the diffusion weighted MR signal. NMR in Biomedicine 23 (7), 661 -- 681. URL http://dx.doi.org/10.1002/nbm.1520 Zhang, H., Hubbard, P. L., Parker, G. J., Alexander, D. C., 2011. Axon diam- eter mapping in the presence of orientation dispersion with diffusion MRI. Neuroimage 56 (3), 1301 -- 15. URL http://www.ncbi.nlm.nih.gov/pubmed/21316474 Zhang, Y., Brady, M., Smith, S., 2001. Segmentation of brain MR im- ages through a hidden markov random field model and the expectation- maximization algorithm. IEEE Trans Med Imaging 20 (1), 45 -- 57. URL http://www.ncbi.nlm.nih.gov/pubmed/11293691 Supplementary Information Section I provides the derivation of the kurtosis in the Ap- pendix A, beyond GPA. Section II provides supplementary data of five subjects. I. Kurtosis of the diffusion on a simple lattice Considering a molecule randomly walking on a one-dimen- sional lattice, we assume that, in each step, the molecule has equal probability to walk to the left and the right. Starting from the origin, after walking N steps, the molecule is away from the origin by n steps. The molecule walks (N + n)/2 steps to the right and (N − n)/2 steps to the left. Therefore, the diffusion propagator is given by where the first 1/2 is a normalization constant such that (cid:80)N n=−N Gn,N = 1 for N (cid:29) 1. Using Stirling's formula for factorials, n! ≈ √ 2πn(cid:0) n 2 , 2 N ! (cid:1)! (cid:1)!(cid:0) N +n (cid:1), the propagator is ap- (cid:18) 1 2 (cid:19)N · (cid:0) N−n (cid:1)n ·(cid:0)1 + 1 12N e Gn,N = · 1 2 proximated by Gn,N ≈ 1√ 2πN − n2 2N − n4 12N 3 , e the correction term exp(cid:0)−n4/12N 3(cid:1). Keeping the lowest or- which is very similar to the propagator of free diffusion except der terms of the correction term, we approximate (cid:18) (cid:19) Gn,N ≈ where C =(cid:0)1 − 1 (cid:82) ∞ 4N 1√ 2πN e− n2 1 − n4 12N 3 2N (cid:1)−1 is a normalization constant such that · C , −∞ Gn,N dn = 1. Using the above propagator Gn,N to calcu- late (cid:104)n4(cid:105) and (cid:104)n2(cid:105), we obtain the kurtosis K ≡ (cid:104)n4(cid:105) (cid:104)n2(cid:105)2 − 3 ≈ − 2 N + O (cid:18) 1 (cid:19) N 2 . II. Supplementary data Fig. S.1 shows scan 1 result in WM ROIs of five subjects. In ACR, SCR, PCR, PLIC, and splenium of the corpus callosum, D⊥ decreases with ∆, manifesting expected ∆-dependence; in contrast, based on the scan 2 result in Fig. S.2, the δ-dependence of D⊥ is too subtle to be individually observed in all WM ROIs. To evaluate the variability between subjects, probability den- sity functions (PDFs) of radial diffusivities of five subjects in WM ROIs are shown in Figs. S.3 and S.4. PDFs of five subjects generally overlap in all WM ROIs, indicating that the variability between subjects is small. 12 Figure S.1: Five subjects' radial diffusivities D⊥ in scan 1 within seven WM ROIs with respect to diffusion time ∆. Figure S.2: Five subjects' radial diffusivities D⊥ in scan 2 within seven WM ROIs with respect to diffusion gradient pulse width δ. 13 0204060801000.30.40.50.60.7ACR0204060801000.30.40.50.60.7SCR0204060801000.30.40.50.60.7PCR0204060801000.30.40.50.60.7PLIC0204060801000.30.40.50.60.7Genu0204060801000.30.40.50.60.7Midbody0204060801000.30.40.50.60.7SpleniumSubject1Subject2Subject3Subject4Subject5010203040500.30.40.50.60.7ACR010203040500.30.40.50.60.7SCR010203040500.30.40.50.60.7PCR010203040500.30.40.50.60.7PLIC010203040500.30.40.50.60.7Genu010203040500.30.40.50.60.7Midbody010203040500.30.40.50.60.7SpleniumSubject1Subject2Subject3Subject4Subject5 Figure S.3: Five subjects' probability density functions (PDFs) of radial diffusivities D⊥ in scan 1 within seven WM ROIs with respect to diffusion time ∆. Figure S.4: Five subjects' probability density functions (PDFs) of radial diffusivities D⊥ in scan 2 within seven WM ROIs with respect to diffusion gradient pulse width δ. 14 -0.500.511.5PDF(ms/µm2)0246∆=26ms,ACR-0.500.511.50246∆=26ms,SCR-0.500.511.50246∆=26ms,PCR-0.500.511.50246∆=26ms,PLIC-0.500.511.50246∆=26ms,Genu-0.500.511.50246∆=26ms,Midbody-0.500.511.50246∆=26ms,Splenium-0.500.511.5PDF(ms/µm2)0246∆=30ms,ACR-0.500.511.50246∆=30ms,SCR-0.500.511.50246∆=30ms,PCR-0.500.511.50246∆=30ms,PLIC-0.500.511.50246∆=30ms,Genu-0.500.511.50246∆=30ms,Midbody-0.500.511.50246∆=30ms,Splenium-0.500.511.5PDF(ms/µm2)0246∆=40ms,ACR-0.500.511.50246∆=40ms,SCR-0.500.511.50246∆=40ms,PCR-0.500.511.50246∆=40ms,PLIC-0.500.511.50246∆=40ms,Genu-0.500.511.50246∆=40ms,Midbody-0.500.511.50246∆=40ms,Splenium-0.500.511.5PDF(ms/µm2)0246∆=55ms,ACR-0.500.511.50246∆=55ms,SCR-0.500.511.50246∆=55ms,PCR-0.500.511.50246∆=55ms,PLIC-0.500.511.50246∆=55ms,Genu-0.500.511.50246∆=55ms,Midbody-0.500.511.50246∆=55ms,Splenium-0.500.511.5PDF(ms/µm2)0246∆=70ms,ACR-0.500.511.50246∆=70ms,SCR-0.500.511.50246∆=70ms,PCR-0.500.511.50246∆=70ms,PLIC-0.500.511.50246∆=70ms,Genu-0.500.511.50246∆=70ms,Midbody-0.500.511.50246∆=70ms,Splenium-0.500.511.5PDF(ms/µm2)0246∆=85ms,ACR-0.500.511.50246∆=85ms,SCR-0.500.511.50246∆=85ms,PCR-0.500.511.50246∆=85ms,PLIC-0.500.511.50246∆=85ms,Genu-0.500.511.50246∆=85ms,Midbody-0.500.511.50246∆=85ms,Splenium-0.500.511.5PDF(ms/µm2)0246∆=100ms,ACR-0.500.511.50246∆=100ms,SCR-0.500.511.50246∆=100ms,PCR-0.500.511.50246∆=100ms,PLIC-0.500.511.50246∆=100ms,Genu-0.500.511.50246∆=100ms,Midbody-0.500.511.50246∆=100ms,Splenium-0.500.511.5PDF(ms/µm2)0246δ=4ms,ACR-0.500.511.50246δ=4ms,SCR-0.500.511.50246δ=4ms,PCR-0.500.511.50246δ=4ms,PLIC-0.500.511.50246δ=4ms,Genu-0.500.511.50246δ=4ms,Midbody-0.500.511.50246δ=4ms,Splenium-0.500.511.5PDF(ms/µm2)0246δ=5ms,ACR-0.500.511.50246δ=5ms,SCR-0.500.511.50246δ=5ms,PCR-0.500.511.50246δ=5ms,PLIC-0.500.511.50246δ=5ms,Genu-0.500.511.50246δ=5ms,Midbody-0.500.511.50246δ=5ms,Splenium-0.500.511.5PDF(ms/µm2)0246δ=6.7ms,ACR-0.500.511.50246δ=6.7ms,SCR-0.500.511.50246δ=6.7ms,PCR-0.500.511.50246δ=6.7ms,PLIC-0.500.511.50246δ=6.7ms,Genu-0.500.511.50246δ=6.7ms,Midbody-0.500.511.50246δ=6.7ms,Splenium-0.500.511.5PDF(ms/µm2)0246δ=10ms,ACR-0.500.511.50246δ=10ms,SCR-0.500.511.50246δ=10ms,PCR-0.500.511.50246δ=10ms,PLIC-0.500.511.50246δ=10ms,Genu-0.500.511.50246δ=10ms,Midbody-0.500.511.50246δ=10ms,Splenium-0.500.511.5PDF(ms/µm2)0246δ=15ms,ACR-0.500.511.50246δ=15ms,SCR-0.500.511.50246δ=15ms,PCR-0.500.511.50246δ=15ms,PLIC-0.500.511.50246δ=15ms,Genu-0.500.511.50246δ=15ms,Midbody-0.500.511.50246δ=15ms,Splenium-0.500.511.5PDF(ms/µm2)0246δ=25ms,ACR-0.500.511.50246δ=25ms,SCR-0.500.511.50246δ=25ms,PCR-0.500.511.50246δ=25ms,PLIC-0.500.511.50246δ=25ms,Genu-0.500.511.50246δ=25ms,Midbody-0.500.511.50246δ=25ms,Splenium-0.500.511.5PDF(ms/µm2)0246δ=45ms,ACR-0.500.511.50246δ=45ms,SCR-0.500.511.50246δ=45ms,PCR-0.500.511.50246δ=45ms,PLIC-0.500.511.50246δ=45ms,Genu-0.500.511.50246δ=45ms,Midbody-0.500.511.50246δ=45ms,Splenium
1505.04643
2
1505
2015-05-19T08:15:09
Emergence and Persistence of Collective Cell Migration on Small Circular Micropatterns
[ "physics.bio-ph", "q-bio.CB" ]
The spontaneous formation of vortices is a hallmark of collective cellular activity. Here, we study the onset and persistence of coherent angular motion (CAMo) as a function of the number of cells $N$ confined in circular micropatterns. We find that the persistence of CAMo increases with $N$ but exhibits a pronounced discontinuity accompanied by a geometric rearrangement of cells to a configuration containing a central cell. Computer simulations based on a generalized Potts model reproduce the emergence of vortex states and show in agreement with experiment that their stability depends on the interplay of spatial arrangement and internal polarization of neighboring cells. Hence, the distinct migrational states in finite size ensembles reveal significant insight into the local interaction rules guiding collective migration.
physics.bio-ph
physics
Emergence and Persistence of Collective Cell Migration on Small Circular Micropatterns Felix J. Segerer,1, ∗ Florian Thuroff,2, ∗ Alicia Piera Alberola,1 Erwin Frey,2, † and Joachim O. Radler1, ‡ 1Faculty of Physics and Center for NanoScience Ludwig-Maximilians-Universitat Munchen, 2Arnold-Sommerfeld-Center for Theoretical Physics and Center for NanoScience, Faculty of Physics, Ludwig-Maximilians-Universitat Munchen, Theresienstrasse 37, D-80333 Munich, Germany Geschwister-Scholl-Platz 1, D-80539 Munich, Germany The spontaneous formation of vortices is a hallmark of collective cellular activity. Here, we study the onset and persistence of coherent angular motion (CAMo) as a function of the number of cells N confined in circular micropatterns. We find that the persistence of CAMo increases with N but exhibits a pronounced discontinuity accompanied by a geometric rearrangement of cells to a configuration containing a central cell. Computer simulations based on a generalized Potts model reproduce the emergence of vortex states and show in agreement with experiment that their stability depends on the interplay of spatial arrangement and internal polarization of neighboring cells. Hence, the distinct migrational states in finite size ensembles reveal significant insight into the local interaction rules guiding collective migration. PACS numbers: 87.18.Gh,87.17.Jj,87.18.Fx,87.17.Aa The ability of cells to coordinate their motion is es- sential in various biological contexts, notably morpho- genesis [1 -- 3] and tissue repair [4, 5]. In recent studies, monolayers of Madin-Darby canine kidney (MDCK) cells have been investigated as model systems for collective behavior in living systems. Remarkably large scaled cor- relations and swirls in cell migration have been observed and characterized using image correlation and traction force microscopy techniques [5 -- 8]. These emergent pat- terns and correlations were attributed to cell-cell cou- pling, and mechano-transduction mediated by the force- generating cytoskeleton. In fact, dynamic self-ordering into streaming patterns and vortex states appears to be rather generic in assemblies of (self-)propelled objects. They are well known in active systems as diverse as driven biopolymers in motility assays [9 -- 11], bacterial colonies [12 -- 14], and driven granular media [15 -- 17]. Su- perficially, these phenomena may be attributed to a ten- dency of neighboring objects to align their direction of motion, as suggested by flocking models [18]. However, upon closer inspection, there are many important quali- tative differences between all these systems and to date quantitative theoretical models are largely lacking. For cell assemblies, the challenge is that mechani- cal and biochemical interactions between cells as well as internal organization of cells are complex [19], and therefore parameter control is limited. Recent progress in understanding collective behavior of cell assemblies has been fueled by micropatterning techniques which enabled well-controlled in-vitro experimental systems. These techniques have been used to study static adher- ence and intracellular cytoskeleton organization of indi- vidual cells in defined geometries [20]. Importantly, geo- metrical confinement of cells into micropatterned circles has been found to induce persistent rotational motion for systems ranging from two cells [21] to large assem- FIG. 1. (a) Array of MDCK cells seeded on circular mi- cropatterns. (b) Circular patterns occupied by 2-8 cells. Cir- cle size increases in such a way that the average area per cell is constant at approximately 830 µm2. Nuclei are labeled in blue. (c) Schematic of 4 cells rotating within a circular field. blies [22, 23] [24]. There is general consensus that on a macroscopic scale collective cell migration is to a large degree generic and can be explained by different classes of theoretical models including flocking models [18, 25 -- 27], cellular Potts models [28 -- 32], and phase field mod- els [33, 34]. However, the mechanisms underlying the emergence of vortex states are still poorly understood and its relationship to single-cell properties remains un- clear. In particular, a systematic study of the emergence and stability of small-scale vortex states and the dynamic disorder-order transition leading to the emergence of col- lective migration as a function of the number of involved cells has not been carried out so far. Here, we investigated the emergence of collective rota- tional motion in small circular micropatterns as a func- tion of the number of cells (Fig. 1). The physical sys- tem consists of arrays of circular fields containing 2-8 cells. Cell density is kept constant by increasing field size in line with cell number. We found distinct transi- tions between states of disordered motion (DisMo) and states of coherent angular motion (CAMo). Further- http://journals.aps.org/prl/accepted/de075Y22He31ad4fe8966203671b32084ef612abb Accepted for publication in PRL: c(cid:13) 2015 American Physical Society more, the survival time of the coherent state tends to increase with increasing cell number, but shows a pro- nounced drop between 4 and 5 cells, where the geometric cell arrangement changes from a conformation without a cell in the system center to one including a centered cell. Employing a computational model, based on the cellular Potts model (CPM) [28, 29], which we extended to in- corporate internal polarization and cell-to-cell mechano- transduction [35], we reproduced and explained these fea- tures. Thus, the experimentally observed gradual transi- tion with increasing system size from predominantly er- ratic motion of small cell groups to directionally persis- tent migration of larger assemblies is captured by the theory, underlining the role of internal cell polarity in the emergence of collective behavior. Micropatterns of the extracellular matrix protein fi- bronectin separated by PEGylated cell-repelling areas were fabricated using a plasma-induced patterning ap- proach. Parts of a culture dish (Ibidi) were covered with a polydimethylsiloxan (PDMS) template of the desired pattern. Exposed parts were treated with O2-plasma in a plasma cleaner (electronic diener) and overlaid for 30 min with 1 mg/mL PLL(20)-g[3.5]-PEG(2) (SuSoS). After- wards, the template was removed and the whole surface was briefly exposed to a 50 µg/mL solution of fibronectin (Yo Proteins). MDCK cells were seeded on the struc- tured surface and placed in a temperature-controlled en- vironmental chamber on the microscope stage. Arrays of circles were designed with increasing sizes to accommo- date 2-8 cells (Fig.1). To ensure constant cell density of 830 µm2/cell, for each pattern size, only fields containing the appropriate number of cells were selected for analy- sis. Nuclei were stained using Leibowitz L-15 medium (c-c-pro) containing 15 ng/mL Hoechst 33342 (Invitro- gen). Time lapse movies were recorded at a rate of 6 frames/hour over 50 h using an iMIC automated micro- scope (TILL Photonics). Individual nuclei were tracked using in-house image analysis software. A typical array of circular adhesion sites occupied by MDCK cells is shown in Fig. 1(a). Cells exhibit spontaneous collective rotation within the circular areas (Fig. 1(c)). Periods of CAMo are seen to be interrupted by intervals of DisMo, after which rotation in an arbitrary direction is resumed (for movies see [36]). Increasing the system size cell by cell (Fig. 1(b)), we studied collective rotation as a function of cell number. For each cell i, the center of the nucleus was tracked and recorded in polar coordinates, and the individual angular positions ϕi(t) were calculated (Fig. 2(a)) (for a detailed descrip- tion see section S2 [36]). Typical time courses of ϕi(t) for a system of 7 cells are shown in Fig. 2(b). To filter out small fluctuations which result, for example, from displacements of the nucleus with respect to the geomet- ric center of the cell, we calculated the system angular velocity ΩN (t) as the mean over the individual angular velocities of the N -cell system smoothed over a number FIG. 2. (color). (a) False-color fluorescence image of the nu- clei of seven cells within a circular micropattern. For each nucleus i, the angular position ϕi(t) was evaluated with re- spect to the circle center. (b) Angular positions ϕi(t) of each cell (in colors corresponding to the nuclei in (a)) and normal- ized total angular velocity ξ(t). The classification threshold of ξc = 1/4 is indicated by the red dashed line. Periods of DisMo are highlighted by gray shaded areas. (c) Probability distribution of the mean angular velocity ΩN for systems containing 2 to 8 cells. The distributions are fitted by a sin- gle Gaussian (green) and a mixture of two Gaussians (dashed red). The deviation between the two curves reveals a local maximum at ΩN = 0. (d) Log-log plot of the angular MSD of CAMo (green) and DisMo (red) and its error for assemblies consisting of 8 cells. For other cell numbers see Fig. S4 [36]. of frames nf taken in discrete intervals of Tf = 10 min: N(cid:88) t+nf(cid:88) [ϕi(τ ) − ϕi(τ − nf )]. (1) ΩN (t) = 1 N · n2 f · Tf i=1 τ =t We chose nf = 9 as the best trade-off between smoothing of fluctuations and temporal resolution [38]. For all N , the probability distribution P (ΩN ) displays symmetry breaking into clockwise and counterclockwise rotations. Both directionalities are almost equally represented, with a small bias towards clockwise rotation (see Fig. S2 [36]). Similar chiralities have been reported before [22, 39, 40]. To distinguish periods of CAMo from periods of DisMo, we analyzed the probability distribution P (ΩN). It was found to be approximately Gaussian (Fig. 2(c)). The maximum, ¯ΩN , as well as the standard deviation, σN , decreased with increasing cell number, displaying an almost constant coefficient of variation σN / ¯ΩN = 0.74 ± 0.13. At ΩN = 0, P (ΩN) exhibits a weak second maximum, indicating a state of disordered, i.e. state increases with increasing cell number, but exhibits a pronounced discontinuity between systems containing 4 and 5 cells (Fig. 3(b)). To further explore the mechanism underlying the dis- continuity in persistence time, we monitored the spatial arrangement of cells within the pattern. Fig. 4(a) shows the relative positions of cells with respect to a reference cell. In systems containing up to 4 cells, the cells are pre- dominantly arranged in topologically equivalent positions in the outer regions of the circle. In this configuration, cells in the state of CAMo follow each other in a closed circle. As the number of cells increases to 5, the packing geometry changes abruptly to a conformation in which a single cell is located at the system center. To connect this topological transition to the observed decrease in the persistence of the CAMo state, intrinsic cell properties have to be accounted for. It is generally assumed that a migrating cell is highly polarized with respect to protein distribution and cytoskeletal organization [19, 41]. In ad- dition, since neighboring cells are coupled mechanically by cell-cell adhesion, a cell obtains directional guidance cues from adjacent cells. This coupling suggests that ad- jacent cells tend to align their direction of internal front- rear polarization. Hence, a ring-like arrangement, as seen for 2-, 3-, and 4-cell systems, naturally provides a stable conformation during a period of CAMo (Fig. 4(b)). If, however, a cell is located in a central position, as in the case of 5 cells, this cell cannot establish a stable axis of internal polarization. It seems likely that this lack of ori- entation leads to the elevated instability we observed for CAMo states of such systems. To test these heuristic ideas we have developed a com- putational model [35] generalizing the CPM [28, 29] to account for both internal cellular polarization and inter- cellular coupling. In the CPM, a cell is represented as a simply connected set of grid sites on a two-dimensional lattice, and thereby cell shape is explicitly represented. The model accounts for mechanical properties of cells and cell-cell adhesion. Previous generalizations of the CPM have implemented cell polarity and ensuing cell migration in a global fashion [31, 32] upon adapting ideas from flock- ing models [18]: the overall polarity of a cell is described by a polarity vector, and it is assumed that there is a pos- itive feedback between a cell's displacement and polarity. While these assumptions provide a simple and efficient way to model interactions between a cell and its mechan- ical environment, they do not resolve internal polariza- tion mechanisms. In fact, there are complex biochemical networks, including Rho family GTPases and membrane lipids, that regulate the assembly of the actin cytoskele- ton and thereby the formation of cell protrusions. Re- cently, computational models have been developed which couple rather sophisticated reaction-diffusion networks to the dynamics of membrane protrusions [42, 43]. These studies have provided important insights into the spa- tially resolved signaling processes within cells and how (color online). (a) Survival function SN (t) = FIG. 3. PN (T > t) of CAMo and DisMo states. Insets show cor- responding log-lin plots. Exponential fits are indicated by dashed lines (for other cell numbers see Fig. S5 [36]). (b) Persistence time τ as a function of cell number, derived from experiment and theory. Error bars indicate confidence bounds of 99% within the fits. (c) Peak positions ¯ΩN of the distri- bution of the angular velocity P (ΩN ) from experimental data and theory. Error bars indicate the standard deviation. ξN (t) := (cid:12)(cid:12)ΩN (t)/ ¯ΩN (cid:12)(cid:12), we defined a common threshold non-rotating, motion. Introducing a normalized variable for all N at ξc = 1/4, so that for ξN (t) < ξc a migration state is classified as DisMo and for ξN (t) ≥ ξc as CAMo respectively (Fig. S2 [36]). (As discussed in section S3 [36], an alternative approach to identify collective motion gave the same results). To verify that these two states are distinct in their migrational behavior we calculated the angular mean squared displacement (MSD) during each state, MSD(t) = (cid:104)[(cid:104)ϕ(t)(cid:105)N −(cid:104)ϕ(0)(cid:105)N ]2(cid:105)states, where t = 0 signifies the starting point of an interval. Averages were taken over all N cells within a given system as well as over all observed intervals of CAMo or DisMo, denoted by (cid:104). . .(cid:105)N and (cid:104). . .(cid:105)states, respectively. Consistently, the MSD of CAMo shows a slope 2 in a log-log plot, indicat- ing ballistic angular motion for all cell numbers, while the MSD of DisMo exhibited diffusive behavior (Fig. 2(d)). Next we evaluated the lifetimes of the CAMo and DisMo states. Fig. 3(a) shows the survival probability SN (t) = PN (T > t), i.e. the fraction of CAMo/DisMo time periods T exceeding t, based on a sample size of over 600 systems (see Table S2 [36]). We found that the survival probabilities of both states decay exponen- tially, SN (t) ∝ e−t/τ suggesting that the stochastic pro- cess underlying the emergence and collapse of both states is Poissonian. The persistence time τ of the coherent FIG. 4. (color). (a) Heat map of the relative nuclei positions with respect to a reference nucleus located at the lower border of the system from experimental and theoretical data (for details on the plot generation see section S8 [36]). (b) Schematic of possible polarization alignments during CAMo for different cell numbers. (c) For the CPM, mean magnitude of polarization and its error is plotted against the radial cell position r normalized by the maximal radial cell position r0. they are affected by cell shape. Here, in order to describe the dynamics off small cell groups, we used an intermedi- ate approach between flocking-type CPM models for cell assemblies [31, 32] and detailed reaction-diffusion models for individual cells [42, 43]. Specifically, in our compu- tation model we employed an internal polarization field within each individual cell to achieve the spatial resolu- tion of microscopic models. At the same time, the numer- ical algorithm is entirely rule-based (rather than based on complex reaction-diffusion networks) to retain the com- putational efficiency of CPMs. Furthermore, to account for the effects of cell-cell communication via mechano- transduction, the local dynamics of the internal polar- ization field is coupled to a cell's membrane protrusions over a finite signaling range. This creates a positive feed- back loop integrating intracellular fluctuations and ex- ternal (mechanical) stimuli and gives rise to spontaneous cell polarization. To match the rotation statistics to the experiments, we simulated cells of fixed (average) size on circular islands at fixed cell density. We then performed a parameter sampling by varying cell adhesion, the range of intracellular mechanical signaling, and the strength of cy- toskeletal forces relative to contractile forces. For a more detailed and technical description of the model please re- fer to S1 in the supplementary material [36], which also contains a list of the model parameters used. The model reproduces the symmetry breaking into ro- tational states found by experiment (see [36] for movies). Analyzing the numerically generated cell tracks analo- gously to experimental data, we found CAMo as well as DisMo (see section S7 [36]). Monte Carlo Steps were ad- justed to real time by matching the CAMo peak positions ¯ΩN . We found the same steady decrease of ¯ΩN with in- creasing cell number as in the experiments (Fig. 3(c)). Moreover, simulation data also exhibit an increase in CAMo persistence with increasing cell number for 2-,3- and 4-cell systems, and reproduce the discontinuity be- tween 4- and 5-cell systems (Fig. 3(b)). (This feature is also observed when alternative measures for persistence are used; see S6 [36].) The discontinuity in persistence is accompanied by the same topological transition in cell ar- rangement as found by experiment (Fig. 4(a) lower part). Assessing the characteristics of internal cell polarization in the model, we found a systematic decrease of the mean magnitude of the front-rear polarization with decreasing distance to the system center (Fig. 4(c)). These findings clearly show that a cell in the center of the pattern is un- able to establish a stable axis of polarization and hence destabilizes collective behavior throughout the system. Here, we have presented a mesoscopic experimental setup in which the emergence and persistence of col- lective behavior is analyzed for a small number of cells in confined geometry. We showed that it is possible to obtain controlled migrational cell states, which may be classified as disordered and coherent angular motion. Both experiments and simulations showed consistently that persistence of the coherent state increases with the number of confined cells for small cell numbers but then drops abruptly in a system containing 5 cells. This is attributed to a geometric rearrangement of cells to a configuration with a central only weakly polarized cell. It reveals the decisive role of the interplay between local arrangement of neighboring cells and the internal cell polarization in collective migration. Future studies combining well-controlled cell assemblies confined to micropatterns and computational models may help to evaluate and characterize migrational phenotypes and identify mechanisms that play a key role in cell-to-cell mechano-transduction and finally in the emergence of collective behavior. This research was supported by the German Excellence Initiative via the program 'NanoSystems Initiative Mu- nich' (NIM), and the Deutsche Forschungsgemeinschaft (DFG) via project B01 and project B02 within the SFB 1032. ∗ Felix J. Segerer and Florian Thuroff contributed equally to this work. † [email protected][email protected] [1] A. J. Ewald, A. Brenot, M. Duong, B. S. Chan, and Z. Werb, Dev. Cell 14, 570 (2008). [2] A. Vasilyev, Y. Liu, S. Mudumana, S. Mangos, P. Y. Lam, A. Majumdar, J. Zhao, K. L. Poon, I. Kondrychyn, V. Korzh, and I. A. Drummond, PLoS Biol 7, e9 (2009). [3] V. Lecaudey and D. Gilmour, Curr. Opin. Cell. Biol. 18, 102 (2006). [4] T. J. Shaw and P. Martin, J. Cell Sci. 122, 3209 (2009). [5] M. Poujade, E. Grasland-Mongrain, A. Hertzog, J. Jouanneau, P. Chavrier, B. Ladoux, A. Buguin, and P. Silberzan, Proc. Natl. Acad. Sci. U.S.A. 104, 15988 (2007). [6] L. Petitjean, M. Reffay, E. Grasland-Mongrain, M. Pou- jade, B. Ladoux, A. Buguin, and P. Silberzan, Biophys. J. 98, 1790 (2010). [7] T. E. Angelini, E. Hannezo, X. Trepat, J. J. Fredberg, and D. A. Weitz, Phys. Rev. Lett. 104, 168104 (2010). [8] A.-K. Marel, M. Zorn, C. Klingner, R. Wedlich-Soeldner, E. Frey, and J. O. Radler, Biophys. J. 107, 1054 (2014). [9] V. Schaller, C. Weber, C. Semmrich, E. Frey, and A. R. Bausch, Nature 467, 73 (2010). [10] T. Butt, T. Mufti, A. Humayun, P. B. Rosenthal, and J. E. Molloy, J. Biol. Chem. S. Khan, S. Khan, 285, 4964 (2010). [11] Y. Sumino, K. H. Nagai, Y. Shitaka, D. Tanaka, K. Yoshikawa, H. Chat´e, and K. Oiwa, Nature 483, 448 (2012). [12] C. Dombrowski, L. Cisneros, S. Chatkaew, R. E. Gold- stein, and J. O. Kessler, Phys. Rev. Lett. 93, 098103 (2004). [13] E. Ben-Jacob and H. Levine, J. R. Soc. Interface 3, 197 (2006). [14] H. Wioland, F. G. Woodhouse, J. Dunkel, J. O. Kessler, and R. E. Goldstein, Phys. Rev. Lett. 110 (2013). [15] V. Narayan, S. Ramaswamy, and N. Menon, Science 317, 105 (2007). [16] A. Kudrolli, G. Lumay, D. Volfson, and L. S. Tsimring, Phys. Rev. Lett. 100, 058001 (2008). [17] C. A. Weber, T. Hanke, J. Deseigne, S. L´eonard, O. Dau- and H. Chat´e, Phys. Rev. Lett. 110, chot, E. Frey, 208001 (2013). [18] T. Vicsek, A. Czir´ok, E. Ben-Jacob, I. Cohen, and O. Shochet, Phys. Rev. Lett. 75, 1226 (1995). [19] A. J. Ridley, M. A. Schwartz, K. Burridge, R. A. Firtel, M. H. Ginsberg, G. Borisy, J. T. Parsons, and A. R. Horwitz, Science 302, 1704 (2003). [20] M. Th´ery, J. Cell Sci. 123, 4201 (2010). [21] S. Huang, C. P. Brangwynne, K. K. Parker, and D. E. Ingber, Cell Motil. Cytoskeleton 61, 201 (2005). [22] K. Doxzen, S. R. K. Vedula, M. C. Leong, H. Hirata, N. S. Gov, A. J. Kabla, B. Ladoux, and C. T. Lim, Integr. Biol. 5, 1026 (2013). [23] M. Deforet, V. Hakim, H. Yevick, G. Duclos, and P. Sil- berzan, Nat. Commun. 5, 3747 (2014). [24] While Deforet et al. [23] observed global changes in ro- tation direction for wild type MDCK cells on large scale patterns, this was not observed by Doxzen et al. [22]. Since substrate stiffness can influence collective behavior significantly [7, 44], this discrepancy might be due to the different substrates used (PDMS/glass). [25] T. Vicsek and A. Zafeiris, Phys. Rep. 517, 71 (2012). [26] M. Basan, J. Elgeti, E. Hannezo, W.-J. Rappel, and H. Levine, Proc. Natl. Acad. Sci. U.S.A. 110, 2452 (2013). [27] N. Sep´ulveda, L. Petitjean, O. Cochet, E. Grasland- Mongrain, P. Silberzan, and V. Hakim, PLoS Comput. Biol. 9, e1002944 (2013). [28] F. Graner and J. A. Glazier, Phys. Rev. Lett. 69, 2013 (1992). [29] J. A. Glazier and F. Graner, Phys. Rev. E 47, 2128 (1993). [30] B. Szab´o, G. J. Szollosi, B. Gonci, Z. Jur´anyi, D. Selmeczi, and T. Vicsek, Phys. Rev. E 74, 061908 (2006). [31] A. Szab´o, R. Unnep, E. M´ehes, W. O. Twal, W. S. Ar- graves, Y. Cao, and A. Czirok, Phys. Biol. 7, 046007 (2010). [32] A. J. Kabla, J. R. Soc. Interface 9, 3268 (2012). [33] B. A. Camley, Y. Zhang, Y. Zhao, B. Li, E. Ben-Jacob, and W.-J. Rappel, Proc. Natl. Acad. Sci. H. Levine, U.S.A. 111, 14770 (2014). [34] J. Lober, F. Ziebert, and I. S. Aranson, Sci. Rep. 5, 9172 (2015). [35] F. Thuroff, M. Reiter, and E. Frey, (manuscript in prepa- ration). [36] See Supplemental Material at http://journals.aps.org/prl , which includes Ref. [37], for time lapse movies, plots and detailed calculations. [37] J. Buhl, D. J. Sumpter, I. D. Couzin, J. J. Hale, E. De- spland, E. R. Miller, and S. J. Simpson, Science 312, 1402 (2006). [38] Data analysis using nf = 4 or nf = 15 respectively had no significant qualitative influence on the results (data not shown). [39] L. Q. Wan, K. Ronaldson, M. Park, G. Taylor, Y. Zhang, J. M. Gimble, and G. Vunjak-Novakovic, Proc. Natl. Acad. Sci. U.S.A. 108, 12295 (2011). [40] L. N. Vandenberg and M. Levin, Dev Biol 379, 1 (2013). [41] N. Nishiya, W. B. Kiosses, J. Han, and M. H. Ginsberg, Nat. Cell. Biol. 7, 343 (2005). [42] A. F. M. Mar´ee, A. Jilkine, A. Dawes, V. A. Grieneisen, and L. Edelstein-Keshet, Bull. Math. Biol. 68, 1169 (2006). [43] A. F. M. Mar´ee, V. A. Grieneisen, and L. Edelstein- Keshet, PLoS Comput. Biol. 8, e1002402 (2012). [44] M. R. Ng, A. Besser, G. Danuser, and J. S. Brugge, J. Cell Biol. 199, 545 (2012).
1701.06079
1
1701
2017-01-21T20:20:58
Abortive Initiation as a Bottleneck for Transcription in the Early Drosophila Embryo
[ "physics.bio-ph", "q-bio.SC" ]
Gene transcription is a critical step in gene expression. The currently accepted physical model of transcription predicts the existence of a physical limit on the maximal rate of transcription, which does not depend on the transcribed gene. This limit appears as a result of polymerase "traffic jams" forming in the bulk of the 1D DNA chain at high polymerase concentrations. Recent experiments have, for the first time, allowed one to access live gene expression dynamics in the Drosophila fly embryo in vivo under the conditions of heavy polymerase load and test the predictions of the model. Our analysis of the data shows that the maximal rate of transcription is indeed the same for the Hunchback, Snail and Knirps gap genes, and modified gene constructs in nuclear cycles 13, 14, but the experimentally observed value of the maximal transcription rate corresponds to only 40 % of the one predicted by this model. We argue that such a decrease must be due to a slower polymerase elongation rate in the vicinity of the promoter region. This effectively shifts the bottleneck of transcription from the bulk to the promoter region of the gene. We suggest a quantitative explanation of the difference by taking into account abortive transcription initiation. Our calculations based on the independently measured abortive initiation constant in vitro confirm this hypothesis and find quantitative agreement with MS2 fluorescence live imaging data in the early fruit fly embryo. If our explanation is correct, then the transcription rate cannot be increased by replacing "slow codons" in the bulk with synonymous codons, and experimental efforts must be focused on the promoter region instead. This study extends our understanding of transcriptional regulation, re-examines physical constraints on the kinetics of transcription and re-evaluates the validity of the standard TASEP model of transcription.
physics.bio-ph
physics
Abortive Initiation as a Bottleneck for Transcription in the Early Drosophila Embryo Alexander S. Serov∗1, 2, Alexander J. Levine1, and Madhav Mani3 1Department of Chemistry And Biochemistry, University of California, Los Angeles, CA 90095, 2Decision and Bayesian Computation Group, Institut Pasteur, 25 Rue du Docteur Roux, Paris, 3Engineering Sciences and Applied Mathematics, Northwestern University, 2145 Sheridan Road, 75015, France USA Evanston, IL 60208, USA January 24, 2017 Abstract Gene transcription is a critical step in gene expression. The currently accepted physical model of transcription of MacDonald, Gibbs & Pipkin (1969) predicts the ex- istence of a physical limit on the maximal rate of tran- scription, which does not depend on the transcribed gene. This limit appears as a result of polymerase "traffic jams" forming in the bulk of the 1D DNA chain at high poly- merase concentrations. Recent experiments have, for the first time, allowed one to access live gene expression dy- namics in the Drosophila fly embryo in vivo under the conditions of heavy polymerase load and test the predic- tions of the model (Garcia et al., 2013). Our analysis of the data shows that the maximal rate of transcription is indeed the same for the Hunchback, Snail and Knirps gap genes, and modified gene constructs in nuclear cy- cles 13, 14, but the experimentally observed value of the maximal transcription rate corresponds to only 40 % of the one predicted by this model. We argue that such a decrease must be due to a slower polymerase elonga- tion rate in the vicinity of the promoter region. This effectively shifts the bottleneck of transcription from the bulk to the promoter region of the gene. We suggest a quantitative explanation of the difference by taking into account abortive transcription initiation. Our calcula- tions based on the independently measured abortive ini- tiation constant in vitro confirm this hypothesis and find quantitative agreement with MS2 fluorescence live imag- ing data in the early fruit fly embryo. If our explanation is correct, then the transcription rate cannot be increased by replacing "slow codons" in the bulk with synonymous codons, and experimental efforts must be focused on the promoter region instead. This study extends our under- standing of transcriptional regulation, re-examines phys- ∗Corresponding author. E-mail: [email protected] ical constraints on the kinetics of transcription and re- evaluates the validity of the standard totally asymmetric simple exclusion process model of transcription. Several ideas on experimental validation of our hypothesis are also suggested. Keywords: transcription, Drosophila fly embryonic development, pattern formation, k-TASEP, extended particles 1 Introduction Pattern formation in the developing organism relies on precise spatiotemporal control of gene expression Gilbert (2013). Gene expression occurs as the result of a complex dynamical process, the first step of which is transcrip- tion Alberts et al. (2002). During transcription, a multi- protein polymerase complex moves sequentially down a strand of DNA Alberts et al. (2002) and produces an mRNA molecule, which, after various downstream pro- cesses, produces a particular protein. While it is clear in some situations that the observed rate of protein pro- duction is selected by the developmental program, that rate is also subject to physical limits resulting from its molecular mechanisms Bialek (2012). Recently, techniques have been developed that allow us to track mRNA levels over time and across nuclear cycles of a fruit fly early development at the individual cell level Garcia et al. (2013), Bothma et al. (2014, 2015). Consequently, we have have access to at least parts of the fly's developmental program (as monitored via the transcription rates of a handful of specific genes) from the first cell division of the fertilized egg through the formation of a complex and spatially patterned embryo. These data provide a novel window on the unfolding 1 7 1 0 2 n a J 1 2 ] h p - o i b . s c i s y h p [ 1 v 9 7 0 6 0 . 1 0 7 1 : v i X r a of the fly's developmental pattern. It is tempting to in- terpret the spatiotemporal changes in the rate of mRNA transcript production solely in terms of evolutionary op- timization of embryonic development. The development process, however, relies on physical processes, which im- pose their own constraints. In this manuscript, we exam- ine the mRNA production data for three genes, known to be a part of the head-to-tail patterning systems ac- tive during early fly development Gilbert (2013), with an eye towards determining whether the physical dynamics of transcription impose limitations on the developmen- tal dynamics. To put this another way: we ask whether the observed developmental program runs in a dynamical regime where it saturates the physical constraints of the underlying biochemical machinery. There is a distinct reason to imagine that the gene ex- pression machinery does, in fact, saturate some dynam- ical upper bounds. On the one hand, there may exist selection pressure on the speed of embryo growth. On the other hand, there are speed limits to transcription inherent in the mechanism itself. In particular, the gene is a linear array of base pairs that must be read sequen- tially; the polymerase complex is doing the reading by stochastically making steps along the gene, and while do- ing so, it takes up a finite-length footprint on it (Fig. 1). A number of such polymerase complexes may simulta- neously run along the gene Garcia et al. (2013), that in a first approximation can be assumed to interact only through steric repulsion (excluded volume). During max- imal protein production, the number of such complexes is large enough (up to 100, Garcia et al. (2013)) that their combined footprint takes up a significant fraction of the gene's length (up to 80 %, Garcia et al. (2013)). At such high coverages, one might well expect that the steric interactions between stochastically moving poly- merase complexes set a maximal rate for gene expression. These basic dynamics are reminiscent of traffic jams ob- served on congested roadways, and has been well studied in a variety of contexts Schutz (2001), Schmittmann and Zia (1995), as we describe below. Here we use earlier developed models to determine from theoretical considerations alone what that maxi- mal mRNA production rate should be and ask whether this rate is indeed observed in the data. It is reason- able to imagine that Drosophila embryonic development is precisely the right place to look for nature to saturate this limiting value of protein production, since optimal production likely leads to the most efficient method for rapidly making a fly. Our findings are, at first, perplex- ing. We indeed find that for three different genes that are transcribed during development, the rates of tran- scriptional initiation production reach the same maxi- mal value (within 15 % of each other), suggesting the presence of a physically-imposed production speed limit that developmental biology saturates, at least in these extreme circumstances. The observed speed limit, how- ever, is only around 40 % that predicted by basic the- oretical considerations, to be described below. A likely explanation of this result is that we have misidentified the rate-limiting step in the mRNA production. One cannot discount the existence of other bottlenecks in the system, and we note here that abortive initiation of transcription has been independently observed in poly- merase dynamics in the proximity of the promoter in in vitro setups Revyakin et al. (2006), Margeat et al. (2006), Kapanidis et al. (2006). In effect, it was reported that the polymerase complex gets stuck for some time at the start of a gene. When we take into account this aspect of transcription and perform calculations based on an independently measured abortive initiation time constant Revyakin et al. (2006), we find that the traffic jams located at the start of the gene further limit the polymerase throughput and bring the predicted maximal rate inline with the observed one. The remainder of this manuscript is organized as fol- lows. First we briefly review the basic physics of traffic jams in one-dimensional models (see also Schmittmann and Zia, 1995, Zia et al., 2011, Chou et al., 2011). We then introduce a modified model based on our knowl- edge of abortive transcription initiation, and discuss our calculations and simulations. The modified model is a special case of what is known in the literature as an inhomogeneous TASEP model with slow (or defective) sites Zia et al. (2011), Dong et al. (2009), Chou and Lakatos (2004). In the next section we present the data and show that these cannot be satisfactory explained by the original model, but are quantitatively consistent with a transcriptional bottleneck associated with the initia- tion of the transcription process observed in the modi- fied model. In the conclusions we discuss the implication of the demonstrated effect for developmental biology and for the statistical physics of non-equilibrium systems, and propose several directions for future research. 2 The statistical mechanics of traffic jams Recognizing that the collective dynamics of many poly- merases walking on DNA can be thought of as that of processive stochastic walkers with simple steric (ex- cluded volume) interactions, we briefly review this long- standing model of non-equilibrium statistical physics. The physics of this one-dimensional model provides per- haps the simplest example of non-equilibrium steady states in a strongly interacting many-body system. This model, originally proposed by MacDonald and collabo- rators MacDonald et al. (1968), MacDonald and Gibbs (1969) and commonly referred to as a Totally Asymmet- ric Simple Exclusion Process (TASEP), has been stud- ied as part of a general interest in the physics of driven diffusive systems Schmittmann and Zia (1995) and as a model for biologically relevant dynamics Zia et al. (2011), Chowdhury et al. (2005), Frey et al. (2004), with the ap- 2 Figure 1: Schematic representation of the abortive k-TASEP model with extended particles of length (cid:96). The abortive k-TASEP model features an additional slow first site (shown in red) compared to the standard k-TASEP model. From the biological point of view, this model represents sequential elongation of polymerase complexes with footprint (cid:96) along a gene of length L plications to ribosome motion along mRNA. We refer the interested reader to Refs. Schmittmann and Zia (1995), Zia et al. (2011), Chou et al. (2011) for extensive reviews of the literature on TASEP. In brief, it is essential to recall that the basic prob- lem is defined by a particle injection rate α at the start (e.g., the left end of the one dimensional track on which the particles move) and an extraction rate at the right end (Fig. 1). The injection rate α may be interpreted as the attempt frequency to add a particle to the left, which is successful only when that space is unoccupied. The extraction rate β may be similarly interpreted. Once injected, particles move stochastically to the right. We will consider only the case of Poisson walkers taking sin- gle steps on a lattice so that, without interparticle in- teractions, a single stepping rate k determines both the mean velocity of the walker and all its velocity fluctua- tions van Kampen (1992). We also restrict our analysis to the case in which particles cannot be added or re- moved in the bulk (i.e. no Langmuir kinetics), although novel dynamics have been studied in this case Parmeg- giani et al. (2003). This restriction is demanded by the experimentally observed processivity of the polymerase complex on the DNA. The model contains two more con- trol parameters: the length of the track L (i.e., the total number of base pairs in the gene) and the length of the walking particles (cid:96) (i.e., the number of lattice sites taken up by one walker). In most TASEP studies the walkers are assumed to be "point particles" (taking up one site) on a lattice constant in length, but, as we will see be- low, the finite length of the polymerase complex plays a significant role in understanding the collective dynamics of transcription in the developing embryo and cannot be ignored. The main phenomenology of TASEP dynamics of point-like particles can be characterized by a phase di- agram defined by the injection and extraction rates α and β respectively Schutz (2001), Zia et al. (2011). It is remarkable that this one-dimensional system with short- ranged interactions can develop long-range order such that the bulk properties of the system can be controlled by these boundary conditions at the endpoints. This is one of the many features distinguishing non-equilibrium steady states from the better understood thermodynamic phases of equilibrium states. A mean-field analysis of the steady states leads to the prediction of three dis- tinct phases Schutz (2001). For α, β sufficiently small, one finds a stable shock front where the particle den- sity jumps from a bulk value defined by the bulk jump rate k to the one controlled by the boundary extrac- tion/injection rates α, β. The position of this stable shock within the system depends on the relative values of α and β. For sufficiently high values of both α and β, one observes a different phase, the maximal current regime. For α > β (β > α) there is also a high (low) density phase in which the current through the system is below the maximal value. These mean-field results are supported by exact solutions to the steady-state density profile obtained by Liggett Liggett (1975) and by numer- ical simulations. In the case of extended particles (the k-TASEP model), the conclusions of the mean-field approach stay valid, but additional factors appear in now approximate formulas for the mean elongation rate s and particle density ρ. For instance, in the low-density (LD) phase for low in- jection attempt rate α and high exit attempt rate β one obtains MacDonald and Gibbs (1969), Shaw et al. (2003): ρLD(α) = α k + α((cid:96) − 1) , sLD(α) = α(k − α) k + α((cid:96) − 1) . (1) √ When the injection attempt rate α increases to its crit- l), while β stays high (β (cid:62) χ), ical value χ = k/(1 + the system experiences a phase transition to the maximal current phase that is characterized by the maximal par- ticle density ρmax and the maximal particle elongation rate smax: ρmax = 1√ (cid:96)(1 + √ , (cid:96)) smax = √ k (1 + . l)2 (2) For more details on the phase diagram of the k-TASEP model, the reader is referred to Refs Shaw et al. (2003), Schutz (2001). Assuming the footprint of a polymerase complex on the DNA is (cid:96) = 45 ± 5 base pairs (bp) long (as determined by electron crystallography, (?Selby 3 Complete entry rulekβ=kParticle of length ࡁWithout steric hindrance particles jump 1 site at a time with rate kIncremental exit with no steric hindranceJump suppressed if next site is occupiedAbortive transcription initiation makes particles remain on the (cid:31)rst site for an average time τα et al., 1997, Poglitsch et al., 1999)), and its bulk jump rate on the DNA is around k ≈ 26 ± 2 bp/s (Garcia et al., 2013), in the maximal current regime, the theory predicts the maximal polymerase injection rate smax ≈ 0.44±0.05 pol/s ≈ 26±3 pol/min. New polymerase com- plexes cannot be recruited on the given gene faster than this rate, which is defined by how fast the injected parti- cle clears the injection site (the first (cid:96) sites of the lattice). On a gene L = 5400 bp base pairs long (like Hunchback gene, (Garcia et al., 2013)), the maximal number of poly- merases is Nmax ≡ Lρmax ≈ 104 ± 11 pol. 3 Abortive k-TASEP Model In this paper we suggest that abortive initiation of tran- scription can be the rate-limiting step of transcription and should be added to the classical k-TASEP model, when used to explain experimental data in early embryo development. One can expect that in such model the bottleneck of the transcription process will be located in the promoter region at the 5' end of the gene, rather than in its bulk as in the original model. Abortive ini- tiation is an effect wherein polymerase complexes loaded on a promoter region of a gene will several times en- gage in transcription cycles, but will abort and release short RNA products returning to the initial position (Ka- panidis et al., 2006, Revyakin et al., 2006, Margeat et al., 2006). However once a product of 9 to 11 nucleotides is synthesized, the polymerase will break its interactions with the promoter and will proceed with mRNA synthe- sis. It is currently accepted that the transition from the abortive initiation state to processive RNA synthesis oc- curs through a "DNA scrunching" mechanism, when the mechanical stress stored in the "scrunched" DNA dur- ing abortive transcription allows the polymerase to break interactions with the promoter. However other mecha- nisms (namely, "transient excursions", "inchworming") have been proposed as well (Kapanidis et al., 2006). To assess the effect of abortive initiation on the tran- scription rate, we modify the standard k-TASEP model by adding a slow site at the beginning of the lattice. This modification will cause the polymerase to stochas- tically pause at the first site imitating the effect ob- served in the experiment. Experimental studies have revealed that the time spent by a polymerase in the abortive initiation stage in the promoter region is an exponentially-distributed random variable with the aver- age pause time τ ≈ 5± 1 s (Revyakin et al., 2006), which we will use in the simulations below. In what follows, we will call this modification of the k-TASEP model an abortive k-TASEP model. The proposed abortive k-TASEP model is a special case of an inhomogeneous k-TASEP model with only one slow site located at the first site of the lattice. The inhomogeneous TASEP model has attracted significant interest in recent years, and some important analytical and numerical results have been obtained for point par- ticles (see Cook et al., 2013, Poker et al., 2015, Dhi- man and Gupta, 2016, Xiao et al., 2016, and references therein), as well as for extended particles Shaw et al. (2003, 2004b,a), Dong et al. (2007), Klumpp and Hwa (2008), Dong et al. (2009), Zia et al. (2011), Brackley et al. (2011). The standard way to treat a problem with an individual slow site like ours would consist in using mean-field approximations for the sublattices to the left and to the right of the slow site and stitching the two so- lutions together by the flux conservation equation on the slow site Kolomeisky (1998), Shaw et al. (2004a). How- ever, when the only slow site is located at the beginning of the lattice, the problem can be easily reduced to the standard k-TASEP model (a similar problem with point particles was recently considered by Xiao et al. (2016)). Indeed, the slow site can be considered as a boundary condition for the remaining L − 1 sites of the lattice that can be treated as an ordinary (L− 1)-site k-TASEP model. Ignoring entry-rule-related boundary effects, one can then, in the first approximation, define the injection attempt rate α of the rest of the lattice through the pause time τ at the first site: α = 1/τ . For long genes L (cid:29) 1, one can safely assume L−1 ≈ L and using Eq. (1), (2) get the following expressions for the normalized polymerase number in the steady state and normalized polymerase elongation rate: Nss/Nmax = sLD/smax = √ √ (cid:96)(1 + (cid:96)) kτ + (cid:96) − 1 , √ (kτ − 1)(1 + (cid:96))2 kτ (kτ + (cid:96) − 1) (3) . These dependencies are shown in Figs. 2a,b as solid red lines for the biologically relevant range of values of τ = 1–10 s. Note that the normalized polymerase num- ber Nss and the normalized injection rate sLD decrease from Nss/Nmax ≈ 0.74 and s/smax ≈ 0.82 for τ = 1 s to Nss/Nmax ≈ 0.17 and s/smax ≈ 0.19 for τ = 10 s. For the experimentally reported value of τ ≈ 5±1 s Revyakin et al. (2006), the corresponding ratios are Nss/Nmax ≈ 0.30 and s/smax ≈ 0.34. Note that while Nss/Nmax describes the steady-state of the system, the expression for sLD/smax describes the current related to a shock wave of density propa- gating in the system before the steady state is estab- lished. The shock-wave description Schutz (2001) is pos- sible thanks to the fact that the interface separating two distinct phases in the TASEP model is microscopic, i.e. the transition occurs on a scale of several lattice sites Janowsky and Lebowitz (1992, 1994). We verify that the shock-wave approach and the approximate for- mulas (3) ignoring boundary-layer effects correctly de- scribe our abortive k-TASEP model by performing Monte Carlo simulations with the numerical values of parame- ters (cid:96), L and k provided above. We use the "complete entry, incremental exit" rule implying that polymerases 4 (a) (b) (c) Figure 2: Numerical simulations of the maximal polymerase number in the steady state Nss (a), the polymerase injection rate s (b), and the evolution of polymerase number Npol (c) as functions of the mean time τ spent in the abortive initiation regime. Each circle in (a),(b) and each curve in (c) represent the average of 20 simulations; the dashed black line marks the maximal loading rate smax ≈ 26 ± 3 pol/min predicted by the original k-TASEP model (Eq. (2)). Independent experiments report that the mean duration of the abortive initiation phase is around τ ≈ 5 ± 1 s (Revyakin et al., 2006), which gives Nss/Nmax ≈ 0.30 and s/smax ≈ 0.34, or Nss ≈ 31 and s ≈ 8.8 pol/min in the absolute values. Black arrows in (c) indicate the established steady state of the simulation where Nss is evaluated. s corresponds to the slope of the curves in (c), with smax labeled in the plot. Eq. (3) and plots (a) and (b) can be used to assess the mean time τ polymerases spend in the abortive initiation regime based on experimental data for each gene, nc and construct. For instance, for the Hunchback bac construct (Fig. 3), one gets τ ≈ 2.0 s in nc 13 and τ ≈ 2.8 s in nc 14 from the mean Nss, and τ ≈ 2.9 s in nc 13 and τ ≈ 4.3 s in nc 14 from the mean injection rate s (dashed green lines). are loaded on the gene only if the first (cid:96) sites are free, while they leave the gene one site at a time with the bulk rate k (Fig. 1) (Lakatos and Chou, 2003). To fo- cus on the effect of the first slow site, we set the poly- merase injection rate p to the value p = 100 s−1 much greater than the inverse time a polymerase spends on the first site of the lattice p (cid:29) 1/τ for all τ in the range 1–10 s. The slow site is modeled as a stochastic process, in which at each time step ∆t a polymerase can either remain at the first (cid:96) sites or continue transcription with a certain finite probability. The escape probabil- ity pesc = ∆t/τ (the probability to leave the slow site) is calculated based on the chosen values of ∆t and τ . The simulations were performed with the rejection-free kinetic Monte Carlo algorithm (KMC, an event-based algorithm, also known as residence-time or Bortz-Kalos-Lebowitz al- gorithm, see (Bortz et al., 1975)). The theoretical pre- diction (3) is found to perfectly describe simulation re- sults (shown as blue circles in Figs 2a,b), which validates our analytical approach. Fig. 2c completes the simula- tion results illustrating the mean evolution of the system from the non-steady state to the steady state. 4 Comparison to Experimental Data In the present section we compare the predictions of the abortive k-TASEP model to experimental results inferred from MS2 fluorescence data after FISH calibration, as de- scribed in (Garcia et al., 2013). Each data entry (fluores- cence trace) contains the number of RNA polymerase II complexes N (t) currently loaded (transcribing) on a se- lected gene as a function of time t recorded in one nucleus of one Drosophila fly embryo. The whole observation pe- riod is divided into nuclear cycles (nc), at the end of each of which the number of cells in the embryo dou- bles, and all polymerase complexes are unloaded from the gene. We define the time point t = 0 to correspond to the start of nc 10 (Garcia et al., 2013). In this anal- ysis we will consider only the nuclear cycles 13 (starting around t ≈ 35 min) and 14 (starting around t ≈ 50 min), since the genes we analyze are mainly expressed in these nuclear cycles. The following three genes were considered in these study: Hunchback (Hb), Snail (Sn) and Knirps (Kn). All considered genes have one primary and one shadow enhancer that influence the spatio-temporal dynamics of 5 246810 (s)0.20.30.40.50.60.7246810 (s)0.20.30.40.50.60.70.80246t (min)smaxNss020406080 = 1.0 s = 3.0 s = 4.0 s = 5.0 s = 7.0 s = 9.0 s = 10.0 s Table 1: Number of fluorescence traces analyzed for each gene-enhancer construct combination. The number of embryos is given in the parentheses Hunchback Snail Bac No primary No shadow 1965 (11) 979 (9) 1279 (8) 547 (5) 607 (4) 159 (1) Knirps 326 (5) 84 (2) 60 (1) expression of these genes. The data presented below were obtained for each of these genes in three modifi- cations: wild-type (labeled bac), a gene with no primary enhancer (no primary) and a gene with no shadow en- hancer (no shadow ). Table 1 summarizes the number of embryos and fluorescence traces analyzed for each gene- enhancer construct combination. The time resolution of the measurements is 37 s. Figs 3c,d show the distribution of polymerase loading rates s as measured in our experiments for bac Hb in nc 13 and 14. The maximal theoretical polymerase in- jection rate smax ≈ 26± 3 pol/min predicted by the orig- inal k-TASEP model is marked by dashed black lines. One can see that smax appears to envelope all observed experimental values, suggesting that Hb transcription in nc 13 and 14 cannot occur faster than smax. The steady-state polymerase number Nss for the Hunchback bac gene is shown in Figs 3a,b, with the k-TASEP predic- tion Nmax ≈ 104 ± 11 pol shown by a dotted black line. Once again we see that Nmax provides a correct over- all limit for all observed experimental values, suggesting that the original k-TASEP model indeed defines the up- per boundary on the number of polymerase complexes loaded on the gene in nc 13 and 14. Moreover, Figs 4 and 5 demonstrate that the actual mean transcription rate s does not depend on a particular gene, construct or location of the nucleus on the anterior-to-posterior (AP) axis of the embryo, suggesting gene-independent nature of the transcription rate constraint. It is however clear that the mean values of the dis- tributions in Fig. 3 lie well below the predictions of the unmodified k-TASEP model MacDonald and Gibbs (1969), Shaw et al. (2003). The histograms in Fig. 3 clearly demonstrate a downshift from the predicted num- bers, indicating that on average the 1D DNA chain is used significantly lower than its maximum polymerase flow capacity for a homogeneous lattice (with, for in- stance, Nss/Nmax ≈ 0.48 and s/smax ≈ 0.44 for Hb bac, Fig. 3). In this paper we interpret this decrease as that on average no traffic jams are observed in the bulk of the DNA chain for the selected genes and con- structs, and that for them abortive transcription initia- tion is the real bottleneck of transcription (other interpre- tations are also possible, and we return to this discussion later). Our speculations are supported by the abortive k-TASEP model that for independently measured value (a) (b) (c) (d) Figure 3: (a),(b): Probability density function (PDF) of the steady-state number of polymerases Nss loaded on the HunchBack bac gene in nuclear cycles 13 (a) and 14 (b) (all AP positions combined). All data entries are grouped in bins of width 5, and the y-axis is proportional to the number of fluorescence traces falling in the current bin. The theoretically predicted upper limit Nmax ≈ 104 ± 11 pol is shown by a dashed black line. The mean values of the distributions in the figure are: Nss = 56 ± 12 pol (nc 13) and Nss = 46 ± 13 pol (nc 14). (c),(d): PDF of initial polymerase injection rates s for the HunchBack bac enhancer construct in nuclear cycles 13 (c) and 14 (d) (all AP positions combined). All slopes are grouped in bins of width 1, and the y-axis is proportional to the number of fluorescence traces falling in the current bin. The dashed vertical line shows the maximal polymerase injection rate smax ≈ 26 pol/min as predicted by Eq. (2). The mean values of the shown distributions are: s = 13 pol/min (nc 13) and s = 10 pol/min (nc 14). In all figures the total number of traces shown is 682 for nc 13, and 1283 for nc 14. One can see that both Nmax and smax correctly predict the maximal observed values, although the average values are lower. The few outliers above smax and Nss are likely due to noisy experimental data of τ ≈ 5 ± 1 s (Revyakin et al., 2006) predicts a simi- lar decrease: Nss/Nmax ≈ 0.30 and s/smax ≈ 0.34. The slight difference between the predicted and the observed values of Nss/Nmax and s/smax may be (besides other rea- sons such as biological variability) attributed to a slightly different value of τ for Drosophila embryos in vivo com- pared to previous in vitro experiments in Ref. Revyakin et al. (2006). If one assumes the validity of the abortive k-TASEP model, the distributions in Fig. 3 can be ex- plained by the following values of τ : τ ≈ 2.0 s for nc 13 6 050100Max. polymerase number Nss00.010.020.03PDFnc 13050100Max. polymerase number Nss00.0050.010.0150.020.025PDFnc 14050100Max. polymerase number Nss00.010.020.03PDFnc 13050100Max. polymerase number Nss00.0050.010.0150.020.025PDFnc 14 Figure 4: Mean evolution of active polymerase number on the Hunchback (blue), Snail (yellow) and Knirps (red) genes. Each curve corresponds to a specific con- struct (bac, no primary or no shadow) of a particular gene (Hunchback, Snail or Knirps) in one of the nu- clear cycles (13 or 14) and is averaged over the anterior- posterior (AP) position in the embryo. The curves are combined in the same plot to demonstrate that the ini- tial injection rate s does not depend on any of these parameters (some noise is present). The thick black curve shows the average over all the parameters. The dashed straight line shows the maximal theoretical injec- tion rate smax ≈ 26 pol/min as predicted by the standard k-TASEP model (Eq. (2)). The solid straight line shows the initial injection rate s/smax ≈ 0.34 as predicted by the abortive k-TASEP model for τ ≈ 5 s. Eqs (3) can be used to adjust the τ parameter and fit the initial slope more accurately (see main text). All fluorescence traces used throughout this article required synchronization of the beginning of nuclear cycles between different embryos and nuclei. The desynchronization is due to recording started at different moments in different embryos and to the limited temporal resolution of the data. All traces shorter than 10 mins were also removed, since we could not reliably synchronize them with the other traces. The steady state of the system seems to be evolving with time, showing a gradual decrease in the active polymerase num- ber. Under these conditions, we have taken the peak number of polymerases as an estimate of Nss in all pre- sented data Figure 5: Initial injection rate s as a function of the AP position for different constructs (bac, no primary, no shadow) of the Hunchback gene. The error bars show 95 % confidence intervals as estimated by boot- strapping (Efron, 1979). Each curve was slightly shifted sideways to allow all error bars to be visible. The dashed line shows the injection rate predicted by the abortive TASEP model (s/smax ≈ 0.34) for τ = 5 s. This predic- tion fits well the experimental data lying slightly below. Narrow AP windows and low number of data sets avail- able to us for the Snail and Knirps genes (Table 1) did not allow us to draw any reliable conclusions on the de- pendence of s/smax on AP for those genes measured values. It can be seen that experimental distri- butions are about 3 times wider than the calculated ones, which suggests that the noise sources included into the numerical abortive k-TASEP model (namely, stochastic elongation and stochastic abortive initiation duration) are responsible for only about one third of the variabil- ity of the experimental values. Other noise sources (such as stochastic promoter activation, heterogeneous or dy- namic elongation rates, instrumental noise or slope fit- ting) must account for the rest of the width of the distri- butions (for further discussion of noise sources in protein synthesis see (Tkacik et al., 2008)). 5 Discussion and τ ≈ 2.8 s for nc 14 from the polymerase number Nss, and τ ≈ 2.9 s for nc 13 and τ ≈ 4.3 s for nc 14 from the mean injection rate s. These values are shown by dashed green lines in Figs 2a,b. To finalize the comparison between the simulations and the experimental data, it is interesting to compare the widths of the calculated distributions for Nss and s to the widths of experimental distributions in Fig. 3. Fig. 6 shows standard deviations (STD) of Nss/Nmax and s/smax calculated for different values of mean abortive initiation time τ and plotted along the experimentally transcriptional dynamics in the Our analysis of Drosophila fly embryo exhibits a common value of steady- state polymerase number Nss and polymerase injection rate s in the beginning of nuclear cycles 13 and 14 for several analyzed genes and constructs. Although such universal character can be explained by the standard k- TASEP model, the observed values of Nss and s are sig- nificantly lower than the theoretically predicted ones. In the present article we have demonstrated that this dif- ference in slopes can be explained by the abortive ini- tiation process that effectively decreases the polymerase 7 051015Time since the start of a nuc. cyc. (min)0102030405060N0.20.30.40.5AP position00.10.20.30.40.50.60.7s/smaxHbnc 13, bacnc 13, no prnc 13, no shnc 14, bacnc 14, no prnc 14, no sh Figure 6: Relative standard deviation of the steady state number of polymerases (δNss/Nmax) and normalized initial injection slope (δs/smax) calculated numerically as a function of mean abortive transcription initiation duration τ . Experimentally calculated values for the Hunchback bac construct are shown by dashed horizontal lines: δNss/Nmax ≈ 0.115 (nc 13), δNss/Nmax ≈ 125 (nc 14), δs/smax ≈ 0.16 (nc 13), δs/smax ≈ 0.16 (nc 14). The experimental lines show a constant value, since τ dependence was not available from the experiment. Note that the noise level of the experimental data is much higher than the noise predicted by the numerical model, which suggests the existence of additional noise sources in experimental data (see main text) elongation rate in the promoter region of the gene. We have modified the k-TASEP model adding a slow first site, and provided simple formulas and performed Monte Carlo simulations to estimate Nss and s in it. Our re- sults for independently measured abortive transcription initiation duration τ ≈ 5 s give values much closer to the experimental ones, and suggest that the promoter region may be the real bottleneck of the transcription process rather than polymerase "traffic jams" downstream of the gene. It is however important to mention that our hypothe- sis is not the only possible explanation for the observed phenomenon. It seems to be impossible to explain the observed decrease in s/smax and Nss/Nmax without slow sites near the start of the gene, but the distribution of slow sites may take different forms. We have only ana- lyzed the simplest option of only one first slow site. Other options may include several distributed slow sites, clus- ters of slow sites Chou and Lakatos (2004), or inhomo- geneous jump rates (quenched disorder) near the start of the gene Shaw et al. (2004b), Zia et al. (2011), Shaw et al. (2003). Interestingly, the effect of quenched disorder on the particle current in the TASEP model has been shown to mainly reduce to the effect of several slowest sites Shaw et al. (2004b, 2003). Although the physical description of systems with different distributions of slow sites would be similar, the biological processes behind the slow sites may be completely different. For instance, the quenched dis- order model describes the fact that different nucleotides for the building mRNA chain may have different concen- trations in the cell, so that the polymerase complex has to wait longer for some nucleotides than for the others thus modulating the elongation speed. It is interesting to mention here that, for example, for some of the pro- teins of Escherichia coli, the slow codons were reported to be preferentially located within the first 25 codons Chen and Inouye (1990). Regardless of the exact distribution of slow sites, an important practical conclusion of this study is that if the bottleneck of transcription is indeed located in the promoter region, the overall transcription rate cannot be increased by modifications in the bulk of the gene, such as replacement of slow codons with syn- onymous ones Shaw et al. (2004b), Zia et al. (2011), Shaw et al. (2003). Experimental research of higher transcrip- tion rates must then focus on acceleration of the initi- ation dynamic or on substitution of slow codons in the promoter region of the gene. From the physical point of view, it seems to be impos- sible to distinguish different configurations of slow sites having only access to the steady-state dynamics of the experimental system. However, one can probably search for the answer in the non-steady state dynamics, or the slow evolution of the steady-states seen in Fig. 4. For in- stance, if slow sites are distributed along the whole gene, the propagating density wave must slow down each time a new, slower site is discovered. Surprisingly, the curves in Fig. 4 don't seem to exhibit this gradual or stepwise slow-down, suggesting again that the slowest sites are lo- cated near the promoter region. This feature, as well as a local decrease of transcription speed around 1–2 mins into nc 13 (Fig. 4, middle pannel), or the gradual de- crease of polymerase density after the maximum has been reached (Fig. 4), cannot be explained by the abortive k- TASEP model, and require further investigation. 8 24681000.050.10.150.2SimulationData nc 13Data nc 1424681000.020.040.060.080.10.12SimulationData nc 13Data nc 14 From the point of view of molecular biology, one could conceive additional experiments aimed at identification of the role of the abortive transcription process in the decrease of the transcription rate. On the one hand, if one could alter the molecular mechanisms behind the abortive initiation and slow it down, the corresponding decrease in the transcription rate and steady-state poly- merase number could be observed experimentally and compared to the predictions of Eq. (3) for the new val- ues of τ . If on the other hand, the abortive initiation could be accelerated, at some point one can expect that the transcription bottleneck may shift to other involved mechanisms, so that the transcription rate may be de- termined by polymerase "traffic jams" in the bulk of the gene or, for instance, promoter activation kinetics. In either case, the curves in Fig. 4 must reflect the changes. We have also compared the width of the distributions of Nss/Nmax and s/smax in our simulations and in the experimental data. Much wider experimental distribu- tions indicate that the noise sources taken into account in our abortive k-TASEP model cannot satisfactory ex- plain the full variability of experimental data, and that additional noise sources should be investigated in future studies. As an interesting development, one could try to extract additional information about the transcriptional bottlenecks and noise sources of the system by analyz- ing the noise, identifying noise signatures seen in the experimental data and shapes of the experimental dis- tributions (i.e., in Fig. 3) and comparing them to those predicted by the abortive k-TASEP model or other noise sources. Finally, the question of whether the system of poly- merases on the DNA actually reaches a steady non- equilibrium state under physiological conditions of nc 13 and 14 also requires additional investigation. For in- stance, Fig. 4 is inconclusive about whether the tran- scription rate stays constant for some time after reaching the peak value and before starting to decrease. If it is not the case, the validity of the steady-state k-TASEP model as a model of transcription under physiological conditions should be critically re-evaluated. References Alberts, B., Johnson, A., Lewis, J., Raff, M., Roberts, K., and Walter, P., 2002. Molecular Biology of the Cell. Garland Science, New York, 4 edition. Bialek, W., 2012. Biophysics: Searching for Principles. Princeton University Press, Princeton, Oxford. Bortz, A., Kalos, M., and Lebowitz, J. L., 1975. A new algorithm for Monte Carlo simulation of Ising spin systems. J. Comput. Phys., 17(1):10–18. http: //dx.doi.org/10.1016/0021-9991(75)90060-1. of eve stripe 2 expression reveals transcriptional bursts in living Drosophila embryos. Proc. Natl. Acad. Sci. U. S. A., 111(29):10598–603. http://dx.doi.org/ 10.1073/pnas.1410022111. Bothma, J. P., Garcia, H. G., Ng, S., Perry, M. W., Gre- gor, T., and Levine, M., 2015. Enhancer additivity and non-additivity are determined by enhancer strength in the Drosophila embryo. Elife, 4(AUGUST2015):1–14. http://dx.doi.org/10.7554/eLife.07956.001. Brackley, C. A., Romano, M. C., and Thiel, M., 2011. The Dynamics of Supply and Demand in mRNA Trans- lation. PLoS Comput. Biol., 7(10):e1002203. http: //dx.doi.org/10.1371/journal.pcbi.1002203. Chen, G. F. T. and Inouye, M., 1990. Suppression of the negative effect of minor arginine codons on gene expression; preferential usage of minor codons within the first 25 codons of the Escherichia coli genes. Nu- cleic Acids Res., 18(6):1465–1473. http://dx.doi. org/10.1093/nar/18.6.1465. Chou, T. and Lakatos, G., 2004. Clustered bottlenecks in mRNA translation and protein synthesis. Phys. Rev. Lett., 93(19):1–4. http://dx.doi.org/10.1103/ PhysRevLett.93.198101. Chou, T., Mallick, K., and Zia, R. K. P., 2011. Non- equilibrium statistical mechanics: from a paradig- matic model to biological transport. Reports Prog. Phys., 74(11):116601. http://dx.doi.org/10.1088/ 0034-4885/74/11/116601. Chowdhury, D., Schadschneider, A., and Nishinari, K., 2005. Physics of transport and traffic phenomena in biology: From molecular motors and cells to organ- isms. Phys. Life Rev., 2(4):318–352. http://dx.doi. org/10.1016/j.plrev.2005.09.001. Cook, L. J., Dong, J. J., and Lafleur, A., 2013. In- terplay between finite resources and a local defect in an asymmetric simple exclusion process. Phys. Rev. E - Stat. Nonlinear, Soft Matter Phys., 88(4):1–8. http://dx.doi.org/10.1103/PhysRevE.88.042127. Dhiman, I. and Gupta, A. K., 2016. Origin and dynam- ics of a bottleneck-induced shock in a two-channel ex- clusion process. Phys. Lett. Sect. A Gen. At. Solid State Phys., 380(24):2038–2044. http://dx.doi.org/ 10.1016/j.physleta.2016.04.031. Dong, J. J., Schmittmann, B., and Zia, R. K. P., 2007. Inhomogeneous exclusion processes with extended ob- jects: The effect of defect locations. Phys. Rev. E - Stat. Nonlinear, Soft Matter Phys., 76(5):31–33. http: //dx.doi.org/10.1103/PhysRevE.76.051113. Bothma, J. P., Garcia, H. G., Esposito, E., Schlissel, G., Gregor, T., and Levine, M., 2014. Dynamic regulation Dong, J. J., Zia, R. K. P., and Schmittmann, B., 2009. Understanding the edge effect in TASEP with 9 mean-field theoretic approaches. J. Phys. A Math. Theor., 42(1):015002. http://dx.doi.org/10.1088/ 1751-8113/42/1/015002. Efron, B., 1979. Bootstrap Methods: Another Look at the Jackknife. Ann. Stat., 7(1):1–26. http://dx.doi. org/10.1214/aos/1176344552. Frey, E., Parmeggiani, A., and Franosch, T., 2004. Col- lective phenomena in intracellular processes. Genome informatics, 15(1):46–55. http://dx.doi.org/10. 11234/gi1990.15.46. Garcia, H. G., Tikhonov, M., Lin, A., and Gregor, T., 2013. Quantitative imaging of transcription in living Drosophila embryos links polymerase activity to pat- terning. Curr. Biol., 23(21):2140–5. http://dx.doi. org/10.1016/j.cub.2013.08.054. Gilbert, S. F., 2013. Developmental Biology. Sinauer Associates, Inc., Sunderland, MA, 10 edition. Janowsky, S. A. and Lebowitz, J. L., 1994. Exact re- sults for the asymmetric simple exclusion process with a blockage. http: //dx.doi.org/10.1007/BF02186831. J. Stat. Phys., 77(1-2):35–51. Janowsky, S. A. and Lebowitz, J. L., 1992. Finite- size effects and shock fluctuations in the asymmetric simple-exclusion process. Phys. Rev. A, 45(2):618–625. http://dx.doi.org/10.1103/PhysRevA.45.618. Kapanidis, A. N., Margeat, E., Ho, S. O., Kortkhon- jia, E., Weiss, S., and Ebright, R. H., 2006. Initial transcription by RNA polymerase proceeds through a DNA-scrunching mechanism. Science (80-. )., 314(5802):1144–1147. http://dx.doi.org/10.1126/ science.1131399. Klumpp, S. and Hwa, T., 2008. Stochasticity and traffic jams in the transcription of ribosomal RNA: Intriguing role of termination and antitermination. Proc. Natl. Acad. Sci. U. S. A., 105(47):18159–18164. http:// dx.doi.org/10.1073/pnas.0806084105. Kolomeisky, A. B., 1998. Asymmetric simple exclu- sion model with local J. Phys. A. Math. Gen., 31(4):1153–1164. http://dx.doi.org/ 10.1088/0305-4470/31/4/006. inhomogeneity. Lakatos, G. and Chou, T., 2003. Totally asymmetric exclusion processes with particles of arbitrary size. J. Phys. A. Math. Gen., 36(8):2027–2041. http://dx. doi.org/10.1088/0305-4470/36/8/302. Liggett, T. M., 1975. Ergodic theorems for the asymmetric simple exclusion process. Trans. Am. Math. Soc., 213:237–61. http://dx.doi.org/10. 1090/S0002-9947-1975-0410986-7. MacDonald, C. T. and Gibbs, J. H., 1969. Concerning the kinetics of polypeptide synthesis on polyribosomes. Biopolymers, 7(5):707–725. http://dx.doi.org/10. 1002/bip.1969.360070508. MacDonald, C. T., Gibbs, J. H., and Pipkin, A. C., 1968. Kinetics of biopolymerization on nucleic acid templates. Biopolymers, 6(1):1–25. http://dx.doi. org/10.1002/bip.1968.360060102. Margeat, E., Kapanidis, A. N., Tinnefeld, P., Wang, Y., Mukhopadhyay, J., Ebright, R. H., and Weiss, S., 2006. Direct observation of abortive initiation and promoter escape within single immobilized transcrip- tion complexes. Biophys. J., 90(4):1419–1431. http: //dx.doi.org/10.1529/biophysj.105.069252. Parmeggiani, A., Franosch, T., and Frey, E., 2003. Phase coexistence in driven one-dimensional transport. Phys. Rev. Lett., 90(February):086601. http://dx.doi. org/10.1103/PhysRevLett.90.086601. Poglitsch, C. L., Meredith, G. D., Gnatt, a. L., Jensen, G. J., Chang, W. H., Fu, J., and Korn- berg, R. D., 1999. Electron crystal structure of an RNA polymerase II transcription elongation complex. Cell, 98(6):791–798. http://dx.doi.org/S0092-8674(00) 81513-5[pii]. Poker, G., Margaliot, M., and Tuller, T., 2015. Sen- Sci. Rep., 5:12795. sitivity of mRNA Translation. http://dx.doi.org/10.1038/srep12795. Revyakin, A., Liu, C., Ebright, R. H., and Strick, T. R., 2006. Abortive Initiation and Productive Initiation by RNA Polymerase Involve DNA Scrunching. Science (80-. )., 314(5802):1139–43. http://dx.doi.org/10. 1126/science.1131398. Schmittmann, B. and Zia, R. Statistical Mechan- In Domb, C. ics of Driven Diffusive Systems. and Lebowitz, J. L., editors, Phase Transitions Crit. Phenomena. Vol. 17, pages 3–214. Academic Press, London, 1995. http://dx.doi.org/10.1016/ S1062-7901(06)80014-5. Schutz, G. M. Exactly Solvable Models for Many- In Domb, C. Body Systems Far from Equilibrium. and Lebowitz, J. L., editors, Phase Transitions Crit. Phenom., chapter 1, pages 3–255. Academic Press, San Diego, 2001. http://dx.doi.org/10.1016/ S1062-7901(01)80015-X. Selby, C. P., Drapkin, R., Reinberg, D., and Sancar, A., 1997. RNA polymerase II stalled at a thymine dimer: footprint and effect on excision repair. Nucleic Acids Res., 25(4):787–793. http://dx.doi.org/10.1093/ nar/25.4.787. 10 Shaw, L. B., Zia, R. K. P., and Lee, K. H., 2003. Totally asymmetric exclusion process with extended objects: A model for protein synthesis. Phys. Rev. E, 68(2): 021910. http://dx.doi.org/10.1103/PhysRevE.68. 021910. Shaw, L. B., Kolomeisky, A. B., and Lee, K. H., 2004a. Local inhomogeneity in asymmetric simple exclusion processes with extended objects. J. Phys. A, 37(6):2105–2113. http://dx.doi.org/10.1088/ 0305-4470/37/6/010. Shaw, L. B., Sethna, J. P., and Lee, K. H., 2004b. Mean- field approaches to the totally asymmetric exclusion process with quenched disorder and large particles. Phys. Rev. E - Stat. Nonlinear, Soft Matter Phys., 70 (2 1):1–7. http://dx.doi.org/10.1103/PhysRevE. 70.021901. Tkacik, G., Gregor, T., and Bialek, W., 2008. The role of input noise in transcriptional regulation. PLoS One, 3(7):e2774. http://dx.doi.org/10.1371/journal. pone.0002774. van Kampen, N., 1992. Stochastic Processes in Physics and Chemistry. North-Holland Personal Library, Am- sterdam. Xiao, S., Chen, X., and Liu, Y., 2016. Totally asymmet- ric simple exclusion process with a single defect site on boundaries. Int. J. Mod. Phys. B, 30(14):1650083. http://dx.doi.org/10.1142/S0217979216500831. Zia, R. K. P., Dong, J. J., and Schmittmann, B., 2011. Modeling Translation in Protein Synthesis with TASEP: A Tutorial and Recent Developments. J. Stat. Phys., 144(2):405–428. http://dx.doi.org/10. 1007/s10955-011-0183-1. 11
1010.5929
1
1010
2010-10-28T11:59:56
Sub-wavelength surface IR imaging of soft-condensed matter
[ "physics.bio-ph", "cond-mat.soft", "physics.ins-det" ]
Outlined here is a technique for sub-wavelength infrared surface imaging performed using a phase matched optical parametric oscillator laser and an atomic force microscope as the detection mechanism. The technique uses a novel surface excitation illumination approach to perform simultaneously chemical mapping and AFM topography imaging with an image resolution of 200 nm. This method was demonstrated by imaging polystyrene micro-structures.
physics.bio-ph
physics
Sub-wavelength surface IR imaging of soft-condensed matter 1 James H. Rice,1,3 Graeme A. Hill,1 Stephen R. Meech,1 Paulina Kuo,2 Konstantin Vodopyanov,2 Michael Reading1 1School of Chemical Science and Pharmacy, University of East Anglia, Norwich NR4 7TJ, UK. 2Ginzton Laboratory, Stanford University, Stanford, CA 94305-4088, USA. 3 present address, School of Physics, University College Dublin, Belfield, Dublin, Ireland Abstract Outlined here is a technique for sub-wavelength infrared surface imaging performed using a phase matched optical parametric oscillator laser and an atomic force microscope as the detection mechanism. The technique uses a novel surface excitation illumination approach to perform simultaneously chemical mapping and AFM topography imaging with an image resolution of 200 nm. This method was demonstrated by imaging polystyrene micro-structures. 1. Introduction 2 Infrared spectroscopy (IR spectroscopy) is a powerful and well developed method to characterise the chemical composition of materials. IR spectroscopy utilises infrared radiation commonly from a black body emitting source to tune across vibrational absorption frequencies [1]. IR is frequently applied to characterise molecular species by utilising the characteristic vibrational modes (eigenmodes) that each molecular species possess. Microscopy based methods have made great progress in developing techniques that can image on the nanoscale, these approaches can be divided into two main classes, near-field and far-field techniques. IR Imaging using far-field methods is limited by diffraction to length scales of the order of 1-5 micrometers for IR microscopy [2]. It is noted that fluorescence based imaging has successfully over come the diffraction limit in the far-field to enable an image resolution as good as tens of nanometres to be archived [3], however to date no progress has been reported for far-field IR based imaging. Scanning probe techniques such as Atomic force Microscopy (AFM) enable topographic imaging of surfaces with very high spatial resolution, i.e. < 10 nm [4]. Such an approach to imaging is a well developed near-field method. Combining IR spectroscopy with AFM potentially allows for a method that can chemically map materials with nanometre spatial resolution. Such an approach based on combining AFM and IR spectroscopy has been developed [5-18]. 3 IR spectra of analytes and thin films have been recorded by combining a Fourier transform infra-red spectroscopy instrument with an AFM [5,6]. This approach is based on an opto-thermal method that utilizes commercial AFM probes as temperature sensors which enabled measurements of opto-thermal signals induced by absorption of IR radiation [7,8]. This method enabled measurements of samples down to 5 µms in size. An alternative approach to opto-thermal based imaging, referred to as AFMIR, has enabled IR imaging of materials with sub-wavelength lateral resolution [9-12]. This method measures IR absorption directly via measuring local transient deformation in the AFM cantilever induced by an infrared pulsed laser tuned at a vibrational absorbing wavelength generating IR absorption spectral information. AFMIR has been applied to image E. Coli bacteria at N-H resonance frequencies, expitaxial quantum dots in resonance with intra-sub band transitions and live cells monitoring the glycogen band centred at 1080 cm-1 with a sub-diffraction limited image resolution of 60-100 nm [10-12]. This experimental methodology requires the use of an attenuated total internal reflection (ATR) arrangement in combination with IR cyclotron radiation. This has effectively limited the application of the technique to samples of thicknesses on the order of the excitation wavelength due to the requirement of evanescent wave propagation through the sample. In order to study samples which are thicker alternative experimental designs are required. Here, we outline an experimental method for sub-micron IR surface imaging (referred to as s- AFMIR) that directly excites samples surface via a top-down excitation arrangement [13]. Additional novel aspects of the outlined experimental approach include the use of an optical parameter oscillator (OPO) tuneable IR source, replacing cyclotron and black body radiation filaments as sources for IR radiation. 4 This approach is different from other work on work combining AFM and optics to enable sub-diffraction IR imaging. Hillenbrand et al applied a method referred to as scattering-type scanning near-field optical microscopy (s-SNOM) to enable sub- diffraction IR imaging [14-16]. In s-SNOM the metalized tip of an Atomic Force Microscope (AFM) is illuminated by laser light. The effect of this is to excite localized surface phonon-polaritons at the tip apex by the concentrated optical field. This produces a near-field resonance at a material specific frequency close to the LO phonon frequency. This method requires the AFM tip to oscillate and the scattered light is collected in the far-field. Aigouy et al also has applied a oscillating tip and far field light collection methodology to enable sub-diffraction optical imaging [17-18]. The outlined method here contrasts with these approaches to perform sub-diffraction imaging by using the excitation light field to induce a topographic change in the sample, the intensity of which is measured as a function of wavelength in in order to construct an IR spectrum. This requires that the AFM tip be keep in contact with the sample and there is no need to collect the scattered light from the sample. This enables the measurement directly of light absorption. The scattering experiment provides less direct access to the IR absorption spectra of materials as the measured response is a convolution of this and the local dielectric properties. 2. Instrumentation The s-AFMIR experimental set-up was based on an in-house built IR laser source and a commercial Atomic force microscope. Mid-IR radiation was generated using a periodically poled LiNbO3 crystal emitting tuneable IR laser radiation, tuneable over 5 3.13 to 3.57 µm as demonstrated by Vodopyanov [19]. The OPO laser provides a high power infrared light source (compared to black-body sources). The OPO was constructed using an PPLN crystal (Super Optronics Inc) and a nanosecond Nd:YAG pump laser (model NL202, EKSPLA). The PPLN crystal possessing a period of 29.0- 30.6 µm fanning and was kept at a constant temperature of 90 0C using a crystal oven heater (temperature controller, model TC038 and model III crystal heater, HC Photonics Corp). The OPO set up is outlined in Fig 1. Selective filtering of laser radiation was performed using a long pass filter (L.P-2500 nm, Spectrogon). Wavelength output was monitored using the frequency mixed component generated at the 800 nm region of the visible spectrum using an ocean optics fibre coupled CCD and spectrograph. The PPLN crystal was mounted onto a linear stage (Melles Griot) and was scanned with the wavelength output monitored. The output power was monitored to be c.a. 3 mW with pump power of c.a. 400 mW. The IR laser is directed onto the surface of the sample been probed by the AFM tip using mirrors as outlined in Fig 1. The excitation beam was focused using a convex mirror. Careful positioning of the excitation beam with respect to the tip was required to optimise the sample. The AFM (Veeco Explorer system) was used with a scanner with lateral and vertical scans of 100 x 100 µm and 10 µm respectively. A Stanford SR650 programmable filter was used to amplify and filter the signal from the AFM. The filtered signal was routed to the input of an Agilent DSO5012A oscilloscope where it was retrieved using in-house software and analyzed with the WSxM program [20]. Tips were silicon nitride v-shaped cantilever tips (used as received from Veeco). The AFM is operated in contact mode, with a typical setpoint of the tip between 5 and 10 nN. Through analysis of the force curve of the contact and by establishing the setpoint, the AFM can remain in a linear range in contact with the surface in line with 6 the literature [10]. Samples were prepared on CaF2 (Aldrich) or Mica substrates. Analytical grade glycerol (Aldrich), polystyrene (PS) films (Aldrich) and 3 µm sized beads (Aldrich) were used. Polystyrene beads (PSB) were dispersed onto a mica surface and heated. This enabled the beads to be fixed to the substrate and are not free to be moved when interacting with the probing AFM tip. PS film was mounted directly onto the substrate. The PSB (on average 3 µm in diameter) were purchased from Aldrich. A commercial Attenuated Internal reflection infrared (ATR-IR) spectrometer was used to record a reference IR spectrum of polystyrene. 3. Origin of s-AFMIR signal The origin of the s-AFMIR comes from optically induced effects that are directly proportional to IR absorption. The AFM cantilever tip was positioned over the sample with the tip in contact with the sample surface (as outlined schematically in Fig 2). Following the application of the IR radiation source the cantilever tip response was monitored. The specific response studied was a change in the vibrational motion of the tip arising from absorption of radiation by the sample. The energy absorbed is dissipated through thermal and acoustic mechanisms. Propagating acoustic waves create a deformation in the surface topography which can be detected by the AFM tip (see Fig 2). As an IR laser excitation source is tuned into resonance with a vibration mode, absorption of IR radiation increases and in turn the energy absorbed increases. Recording the cantilever displacement intensity as a function of excitation wavelength 7 enables either IR spectra or IR imaging to be performed when the AFM cantilever is stationary or is scanning respectively. The absorption of the excitation radiation exponentially falls as the excitation radiation propagates through the absorbing material [21,22]. As a consequence the applied energy from the excitation radiation will be converted into thermal and acoustic energy predominantly at and near the surface of the sample. The effective spatial resolution been a function of the sample under study. AFMIR has been applied to study quantum dots with a spatial resolution of 60 nm [11]. This indicates that the propagation lengths of the acoustic waves that are responsible for the absorption measurement mechanism are short enough in such materials to enable nanoscale imaging. 4. Studies of bulk material Polystyrene films (20mm x 20 mm x 1 mm) were studied using s-AFMIR. The AFM cantilever tip was positioned over the centre of the PS film with the tip in contact with the sample. Following the application of the IR radiation source the cantilever tip response was monitored. Fig 3 shows an s-AFMIR spectrum recorded for a PS film, shown along with an IR spectrum recorded using a commercial IR instrument. Inspection of the spectra show that both possess the same spectral features. This indicates that the s-AFMIR method has recovered the IR spectrum for a PS film with the characteristic vibration intensities arising from the aromatic ring, (C6H5), giving rise to the group of bands above 3010 cm-1 and bands below 3010 cm-1 present arising from the saturated main chain CH groups. 5. Mesoscopic systems 5.1. Localised spectroscopy 8 Studies of individual PSBs were undertaken to demonstrate imaging small microscopic objects using this approach. A single PSB, 3.5 µm in diameter and 750 nm in height was located using AFM topography scanning. The AFM cantilever tip was positioned over the centre of the PSB with the tip in contact with the sample. Recording an s-AFMIR signal when the AFM tip is stationary enables an IR spectrum of a sample to be recorded as shown in Fig 3. The spectrum is shown along with a reference IR spectrum of a polystyrene film, recorded using a commercial IR instrument for comparison. The s-AFMIR spectra shown in Fig 3 possess a wavelength resolution of 12 cm-1, while the ATR-IR based spectrum has a wavelength resolution of 1 cm-1. The s- AFMIR resolution was due to the scanning increments. It is noted that the limiting resolution will be the PPLN IR source which is on the order of <6 cm-1 with the outlined experimental set up. In order to access this resolution the use of a precision controlled motorised stage or manual stage with micron resolution is required to scan across the PPLN crystal across the Nd:YAG pump laser beam. This must be coupled with precisely focused laser pump excitation beam that matches the resolution in the fanning gradient in the PPLN crystal, noting that the pump power must not exceed the damage threshold of the PPLN crystal. Analysis of the s-AFMIR spectrum of PSB in Fig 3 shows that there are differences in the spectral features when compared to the IR spectrum of the PS film. The band at 2960 cm-1 is larger in relative intensity in the PSB spectrum compared to the PS film. 9 This spectral feature difference may arise from the local structural form that each of the polymers possess. Studies of PS with different morphologies has shown that the 2960 cm-1 band is sensitive to the presence of these morphology differences i.e. spherical beads or linear films of PS, it is noted that the heating of the PSB beads may also result in the formation of different polymer phases resulting in the observed relative intensity changes [23,24]. The AFMIR spectrum was recorded using the FFT of the cantilever oscillation as a function of excitation wavelength. The FFT of the signal shows the presence of numerous peaks corresponding to the different resonant vibrational modes of the AFM tip as a function of frequency. By monitoring the intensity of the FFT peaks it is possible to track the resonant response of the cantilever that is influenced by the absorption behaviour of the sample. A previous report by Hill et al outlined that s- AFMIR imaging showed that through the application of FFT signal processing the signal to noise ratio can be increased [13]. The use of FFT signal processing creates a negligible background for the off-resonance signal making top-down s-AFMIR and ATR AFMIR methods comparable in regard to reduction in off-resonance background signal which demonstrates the high contrast of spectroscopic measurement capabilities in both approaches. The image resolution of s-AFMIR was estimated to be 200 nm. Measuring a line scan across the edge of a PSB and comparing the s-AFMIR resolution with an AFM topography scan recorded at the same time estimated the s-AFMIR resolution to be 200 nm comparable with that of the AFM (see Fig 4). As Fig 4 and Fig 4’s insert at a spatial separation of 200 nm the AFMIR signal be differentiated from each proceeding point producing a line-scan plot that was almost identical to that of the 10 AFM. 5.2. Sub-wavelength imaging Fig 5 shows IR images of a single PBS. Shown is an AFM topography image and IR images at two different excitation wavelengths. The IR images were recorded by monitoring the FFT signal as a function of excitation wavelength when the AFM tip was been scanned. AFM topography images were recorded simultaneously. The IR image was constructed by recording an s-AFMIR spectrum per image point thus building an IR adsorption image. Fig 5 shows an s-AFMIR image recorded of the PSB at an on-resonance excitation wavelength. The white pattern is assigned to arise from IR absorption originating from a strong opto-acoustic oscillation response by the PSB, an absorption which is absent for the mica substrate. Comparing the AFM topography and s-AFMIR images shows good agreement in the image features. Fig 5 also shows an s-AFMIR image recorded of the PSB at an off-resonance excitation wavelength. The s-AFMIR shows a reduced white background in line with reduced IR absorption. 3D plots of s-AFMIR images of a single polystyrene bead are shown in Fig 5. 3D plots enable the analysis of the different IR absorption intensity for the s- AFMIR images on and off resonance to be clearer. 3D plot of an on-resonance s- AFMIR image at 2920 cm-1 is shown along with a 3D plot of off-resonance excitation recorded at 2065 cm-1. Inspection of the two 3D plots shows clear differences in signal intensity. The Plots recreate the physical dimensions of the polystyrene bead as measured with AFM topography methodology with the greater polystyrene s-AFMIR signal coming from the centre of the bead, where AFM topography measurements 11 show the greatest polystyrene mass to be present. Differences between AFM topography and s-AFMIR images which are present may arise from the presence of thermal drift of the mechanical device occurring during the duration of the measurement of s-AFMIR images i.e. 20 min. Analysis of the IR images shows that the left hand side of the PSB is more pronounced than the right hand side. This is potentially associated with the angle of incidence of the excitation laser. Previous AFMIR studies of E.coli [10] reported that the resultant IR images of the bacterium possessed weak intensity at centre of the bacterium and relatively intense signal at the edges. This was explained due to the bacterium scattering the incident electric field around the bacterium resulting in a lightening distribution and not a homogenous distribution. In additional it was noted that the signal is greater at one side than the other which was explained by referring to the angle of incidence of the excitation light. In order to overcome these inhomogeneous distributions requires careful choice of the angle of incidence of the excitation radiation beam. Never the less in the reported case, and in the images presented here, the IR images match very well with the AFM topography images. Analyses of smaller sample sizes were undertaken. Fig 5 shows an AFM topography image and an s-AFMIR image recorded on-resonance at 2965 cm-1 of moulded PSB formed by heating the sample. The AFM topography image shows a maximum height of 367 nm for the sample. The s-AFMIR image replicated the AFM topography features and demonstrates that thin surface features on the order of >367 nm can be imaged. Fig 6 shows the corresponding IR absorption images of the same bead recoded at five different wavelengths. These images were recorded at a 5x times faster image rate than images shown in Fig 5. The IR absorption image recorded on-resonance show an 12 intense white pattern present on a dark background. The white pattern is assigned to arise from IR absorption originating from a strong opto-acoustic oscillation response by the PSB, an absorption which is absent for the mica substrate. The IR absorption image recorded after moving the excitation wavelength off resonance shows the presence of a much weaker white pattern. This is expected as the IR absorption is much weaker but still present, resulting is a weaker but still visible white pattern. Moving further of resonance results in an image with no white pattern visible. While the rapid accumulation of the images results in distortion of the s-AFMIR image compared to those in Fig 5, it allows for relatively rapid analysis of the presence of polystyrene. Inspection of Fig 6 shows a correlation of peak intensity in the localised s-AFMIR spectrum with intensity of the presence of the PSB, as denoted by the intensity of the light contrast colouring. The strongest contrast is seen at the peak of the vibrational band. Imaging on the side band shows an image with a white pattern of intermediate intensity in line with the intensity of the side band seen in the s-AFMIR spectrum shown in Fig 6. 6. Potential improvements Top-down excitation arrangement enables adequate excitation of a surface area of several microns using the out-lined set-up. The focusing of the excitation IR wavelengths via the use of an elliptical mirror practically enables focusing of the IR beam on the order of >10 microns. Focusing at the diffraction limit of λ/2 would enable a beam on the order of two microns to be achieved. This would limit s-AFMIR imaging to length scales on this order as imaging must occur within the IR excitation area. Imaging was performed without tightly focusing the excitation beam in order to excite a large area. This was done at the expense of signal sensitivity which would 13 expect to improve with higher laser excitation power. A number of improvements in the experimental system may lead to a further improvement in image resolution. Studies have shown that the tip motion is sensitive to both tip-surface and the tip-environment interactions [9]. This indicates that choice of tip composition should consider the material system been probed in order to optimise these interactions. This may include optimisation of AFM tip design to lead to more sensitive cantilever response in addition to reducing signal response arising from direct AFM tip excitation by the excitation laser. Studies of the effect of temperature on microcantilever resonance response showed that for mono-material cantilevers frequency varies in direct proportion to the temperature in line with the decrease in Young’s modulus with increasing the temperature. When the cantilever is bi-material, the response is nonlinear due to differential thermal expansion. This indicates the choice of material used to build the tip with is important. Optimisation of AFM tip design may also include introduction of opto-thermal functionalism taking advantage of potential research advances in the development of opto-thermal IR imaging which will increase the range of measurements that can be made using the experimental approach outlined here. Conclusion A technique for sub-wavelength infrared surface imaging and spectroscopy of micro- structured polymeric material has been demonstrated with an image resolution of λ/100. In this way the surface features of a sample can be chemical characterised via 14 inspection of the resulting IR spectra and images. This was achieved by combining a phase matched optical parametric oscillator laser as the IR source and an atomic force microscope as the detection mechanism. The technique uses a novel surface excitation illumination by focusing, at the diffraction limit, the excitation light directly onto the surface of the sample in presence of the AFM tip. This method enables chemical mapping and AFM topography imaging to be performed simultaneously with an image resolution of 200 nm. References 1. B. H. Stuart, Infrared spectroscopy: fundamentals and applications (John Wiley 2004). 2. E. Betzig, J. K. Trautman, T. D. Harris, J. S. Weiner and R. L. Kostelak, Science, 251, 1468 (1991). 3. J. H. Rice, Mol. BioSyst., 3, 781 (2007). 4. T. R. Albrecht and C. F. Quate, J. Vac. Sci. Tech. A., 6, 271 (1988). 5. A. Hammiche, M. H. Pollock, M. Reading, M. Claybourn, P. M. Turner and K. Jewkes, Appl. Spectrosc., 53, 810 (1999). 6. A. Hammiche, L. Bozec, M. J. German, J. M. Chalmers, N. J. Everall, G. Poulter, M. Reading, D. B. Grandy, F. L. Martin and H. M. Pollock, Spectrosc., 19, 20 (2004). 7. A. Hammiche, L. Bozec, H. M. Pollock, M. German and M. Reading, J. Microsc. (Oxf), 213, 129 (2004). 8. M. Reading, D. M. Price, D. B. Grandy, R. M. Smith, L. Bozec, M. Conroy, A. Hammiche and H. M. Pollock, Macromol. Symp., 167, 45 (2001). 9. A. Dazzi, R. Prazeres, F. Glotin and J. M. Ortega, Infr. Phys. Tech., 49, 113 15 (2006). 10. A. Dazzi, R. Prazeres, F. Glotin and J. M. Ortega, Ultramicroscopy, 107, 1194 (2007). 11. J. Houel, S. Sauvage, P. Boucaud, A. Dazzi, R. Prazeres, F. Glotin, J.M. Ortega, A. Miard, A. Lemaıtre, Phys. Rev. Lett., 99, 217404 (2007). 12. C. Mayet, A. Dazzi, R. Prazeres, F. Allot, F. Glotin, and J. M. Ortega, Optics Lett, 33, 1611 (2008). 13. G. Hill, J.H. Rice, S.R. Meech, P. Kuo, K. Vodopyanov, M. Reading, Optics Lett., 34, 431 (2009). 14. J. Aizpurua, T. Taubner, F. J. Garcia de Abajo, M. Brehm , R Hillenbrand, Opt. Expr. 16, 1529 (2008). 15 . Cvitkovic, N. Ocelic, R. Hillenbrand, Nano Lett., 7, 3177 (2007). 16. M. Brehm, T. Taubner, R. Hillenbrand, F. Keilmann, Nano Lett. 6, 1307 (2006). 17. F. Formanek, Y. De Wilde, L. Aigouy, Ultramicroscopy 103, 133 (2005). 18. A. Fragola, L. Aigouy, P.Y. Mignotte, F. Formanek, Y. De Wilde, Ultramicroscopy 101, 47 (2004). 19. K. L. Vodopyanov and P. G. Schunemann, Opt. Lett., 28, 441 (2003). 20. I. Horcas, R. Fernandez, J. M. Gomez-Rodriguez, J. Colchero, J. Gomez-Herrero, A. M. Baro, Rev. Sci. Instrum., 78, 013705 (2007). 21. H. T. Jung, Appl. Phys. B 70, 237 (2000). 22. H. N. Subrahmanyam, S.V Subramanyam, J. Mater., Sci., 22, 2079 (1987) . 23. J. Duchet, S. Demoustier-Champagne, Polymer, 41, 1 (2000). 24. F. Vilaplana, S. Karlsson, A. Ribes-Greus, Euro Poly J., 43, 4371 (2007). a Pump Pump laser laser Nd:YAG selectively coated mi rror PPLN selectively coated mi rror Cube polariser b Exci tation laser lens λλλλ /2 plate Wavelength selector 16 Fig. 1. Schematic drawing of a) periodically poled LiNbO3 crystal emitting tuneable IR laser radiation, b) the experimental top-down excitation alignment. 17 a Can ti lever Absorption of laser pu lse Absorbing Absorbing material material Su rface Su rface deformation deformation Can tilever displacement Deformation Deformation distan ce distan ce b r e v e l i t n a C y t i s n e t n I n o i t a l l i c s o Before excitation Following excitation T ime T ime Fig 2. a) Schematic drawing of the origin of the measured signal, before on-resonance excitation, left, and the resulting surface deformation and cantilever displacement following on-resonance absorption of incident radiation, b) schematic drawing of a dampening oscillation of the cantilever following its response to the surface deformation. 18 ) B d ( l a n g i s R I M F A 1 0 s b A R I T F 0.4 0.1 a PS film ) B d ( l a n g i s R I M F A s b A R I T F 10 b onr PSB offr 0 0.4 0.1 offr onr 60 c 40 20 ) B d ( . n e t n i . t a l l i c s o r e v e l i t n a c 2800 2900 3000 Wavenumbers (cm-1) 2800 2900 3000 Wavenumber (cm-1) 300 100 Freq. (KHz) 500 Fig 3. a) local s-AFMIR spectrum recorded with the AFM tip at a fixed position on a PS film (top), shown with a spectrum of polystyrene recorded using a commercial instrument (bottom), b) local s-AFMIR spectrum recorded with the AFM tip at a fixed position on a single PSB (top), shown with a spectrum of polystyrene recorded using a commercial instrument (bottom), c) FFT spectra of spectral plots of the cantilevers oscillations shown in a) on resonance (onr), and off resonance (offr). . 19 280 240 200 160 O p t o - a c o u s . ( m V ) 400 300 200 100 ) m n ( t h g i e H ) m n ( t h g i e H A F M I R S i g . ( m V 5000 5400 ) Lateral Posit. (nm) 200 nm 200nm 0 4 5 6 7 8 9 Lateral Position (µµµµm) Fig. 4. a) Plot of PTIR (black curve) and AFM topographic(gray curve) data from a single PSB. Inset shows an expanded part of the plot. a 700 nm a 20 b 0 nm 10 dB i 0 dB 10 dB ii 0 dB e d 1 µµµµm 1 µµµµm c b c Excitation direction 0 dB Fig. 5. a) Images recorded from a single PSB, an AFM topography image (top), localised s-AFMIR absorption images recorded at 2920 cm-1 (on resonance) (middle)and an s-AFMIR image at 2965 cm-1 (off resonance) (bottom), shown also are 3D plots of both s-AFMIR images shown along side their corresponding IR images. b) AFM topography (top) image and an s-AFMIR image at 2920 cm-1 (on-resonance) for moulded PSBs. c) arrow shows the IR laser excitation direction using an image from a). 10 ) B d ( . s b A 0 2900 3000 2800 Wavenumber (cm-1) 21 Fig. 6. s-AFMIR Images for a single polystyrene bead recorded at different excitation energies on and around the IR C-H stretch region. Arrow shows the IR laser excitation direction.
1507.01463
1
1507
2015-07-06T13:53:51
Scaling and optimal synergy: Two principles determining microbial growth in complex media
[ "physics.bio-ph", "cond-mat.other", "q-bio.CB", "q-bio.MN" ]
High-throughput experimental techniques and bioinformatics tools make it possible to obtain reconstructions of the metabolism of microbial species. Combined with mathematical frameworks such as flux balance analysis, which assumes that nutrients are used so as to maximize growth, these reconstructions enable us to predict microbial growth. Although such predictions are generally accurate, these approaches do not give insights on how different nutrients are used to produce growth, and thus are difficult to generalize to new media or to different organisms. Here, we propose a systems-level phenomenological model of metabolism inspired by the virial expansion. Our model predicts biomass production given the nutrient uptakes and a reduced set of parameters, which can be easily determined experimentally. To validate our model, we test it against in silico simulations and experimental measurements of growth, and find good agreement. From a biological point of view, our model uncovers the impact that individual nutrients and the synergistic interaction between nutrient pairs have on growth, and suggests that we can understand the growth maximization principle as the optimization of nutrient synergies.
physics.bio-ph
physics
Scaling and optimal synergy: Two principles determining microbial growth in complex media Francesco Alessandro Massucci,1 Roger Guimer`a,2, 1 Lu´ıs A. Nunes Amaral,3, 4, 5 and Marta Sales-Pardo1, ∗ 1Departament d'Enginyeria Qu´ımica, Universitat Rovira i Virgili, 43007 Tarragona, Spain 2Instituci´o Catalana de Recerca i Estudis Avanc¸ats (ICREA), Barcelona 08010, ES 3Department of Chemical and Biological Engineering, Northwestern University, Evanston, IL 60208, USA 4Northwestern Institute on Complex Systems, (NICO) Northwestern University, Evanston, IL 60208, USA 5Howard Hughes Medical Institute, Northwestern University, Evanston, IL 60208, USA (Dated: July 18, 2021) High-throughput experimental techniques and bioinformatics tools make it possible to obtain reconstructions of the metabolism of microbial species. Combined with mathemat- ical frameworks such as flux balance analysis, which assumes that nutrients are used so as to maximize growth, these reconstructions enable us to predict microbial growth. Although such predictions are generally accurate, these approaches do not give insights on how differ- ent nutrients are used to produce growth, and thus are difficult to generalize to new media or to different organisms. Here, we propose a systems-level phenomenological model of metabolism inspired by the virial expansion. Our model predicts biomass production given the nutrient uptakes and a reduced set of parameters, which can be easily determined exper- imentally. To validate our model, we test it against in silico simulations and experimental measurements of growth, and find good agreement. From a biological point of view, our model uncovers the impact that individual nutrients and the synergistic interaction between nutrient pairs have on growth, and suggests that we can understand the growth maximization principle as the optimization of nutrient synergies. 5 1 0 2 l u J 6 ] h p - o i b . s c i s y h p [ 1 v 3 6 4 1 0 . 7 0 5 1 : v i X r a ∗ [email protected] I. INTRODUCTION 2 The rapid development of high-throughput experimental techniques and bioinformatics tools has made it possible to obtain reliable metabolic reconstructions from genomic data in a semi- automatic fashion [1 -- 4]. The availability of such reconstructions makes it possible, in turn, to investigate metabolism from a systems point of view [5]. In particular, the development of a math- ematical framework to predict cellular growth based on cellular function optimization has signif- icantly advanced our understanding of how the metabolic state of an organism will change upon modifications in the growth medium, the introduction of mutations, or the effect of stress [6 -- 12]. Unfortunately, our ability to calculate microbial growth rates has not been paralleled by a sub- stantial gain of insight into metabolic processes, especially for what concerns the impact of nu- trients on growth. A number of mathematical models have been developed aiming at predicting microbial growth rates [13 -- 18], but these models are only valid for a limited number of specific nutrients and are not easily generalizable because of the need to determine parameters empirically. Here, we present a systems-level phenomenological model that enables us to predict growth and, at the same time, provides insights into the effective systems-level principles by which nutri- ents are catabolized. Our approach does not predict which nutrients will be uptaken from a given medium; rather, it predicts, from the values of the uptakes, how each nutrient will contribute to cellular growth. Despite the fact that we use flux balance analysis (FBA) to develop, justify and validate our model (and that, as we discuss later in Section IV, FBA has well known limitations), the model is ultimately independent of FBA and of any particular metabolic reconstruction; in this sense, the model is also organism-independent. Our approach, which is analogous to a virial expansion, reveals that cellular growth can be well-approximated by the contributions of each individual nutrient plus a synergy term that con- siders nutrient-pair contributions. We demonstrate that the predictions of the model are in good agreement with empirical measurements of biomass production. Moreover, our model provides novel insight into the effective contributions to growth since we can express synergy contributions as scaling functions that depend exclusively on four factors: the type of nutrients considered, the pathways that catabolize them, the ratio between their uptake fluxes, and the effective carbon con- tent of each nutrient. Uptake fluxes are allocated among possible synergistic contributions in order to maximize synergy, thus revealing the principles of nutrient use that lead to the maximization of biomass production. II. MODEL Our goal is to express in closed form the steady-state growth rates g of a bacterium given the nutrient uptakes from the external medium, without taking explicitly into account any micro-level information about the processes occurring inside the cell. In [19] and [20], models to predict which nutrients can produce growth and what constraints are necessary to reproduce observed uptakes in 3 rich media were already developed. Here, we consider that the real uptake fluxes of each nutrient are known and fall within the empirical range which ensures that nutrient uptakes can be fully catabolized [9]. To validate our model, we use FBA predictions of biomass production for Escherichia coli using the metabolic reconstruction iAF1260, which has been shown to yield a good agreement with empirically measured growth rates [21]. Note that we focus exclusively on the use of nutrients for biomass polymerization, discarding the role of ATP maintenance (see [19] and Sec. IV). For simplicity, we focus on nutrients that belong to one of the four main nutrient classes: sugars, fatty acids, amino acids, and bases (see Appendix for a complete list). Following a virial expansion-like formulation, we hypothesize that, given a fixed vector of nutrient uptake fluxes φ, we can express the steady-state biomass production of an organism as E(cid:88) (cid:88) i=1 E(cid:88) γ(cid:0)φi, φj, φk j<k αi(φi) + (cid:1) (cid:0)φj, φk (cid:1) + . . . , βjk g(φ) = + (1) where E is the number of uptakes. i<j<k A first order approximation is equivalent to considering that each single nutrient contributes independently to g(φ) as in [19]. In analogy to the ideal gas approximation, we call this model idealized metabolism (IM). Note that because we consider the nutrient use for stationary biomass production exclusively, in the presence of a single nutrient uptake (i.e. φi (cid:54)= 0 for a single i and φk ≡ 0 for k (cid:54)= i) the scale of our system is precisely given by φi. Therefore, the biomass production must be proportional to φi, so that g(φi) = αφi, where α is the biomass yield of nutrient i [9, 19, 22]. For the first order terms, we thus write: g(φ) = αiφi. (2) We evaluate αi for each nutrient i by computing the FBA biomass production gFBA(φ(i)) allowing for a single nutrient uptake i=1 αi(φi) = E(cid:88) E(cid:88) φi = 1 arb.units, φj = 0 arb.units i=1 φ(i) = ∀j (cid:54)= i, where we use arbitrary units, since all fluxes are defined up to a multiplicative constant in the FBA problem. Note that in Eq. (2), only purines among bases can be accounted for growth, since pyrimidines alone cannot be catabolized by E. coli [19]. Previously, we found that αi is proportional to the effective number of carbons Ci, that is, the number of carbons that are actually catabolized 1 in each metabolite i as 1 For the nutrient classes we consider, the effective carbons equal the actual carbons for all nutrients except for the αi = acCi, (3) bases 4 with a slope ac that is nearly insensitive to the nutrient class c (fatty acids, sugars, amino acids, Fig. 1a). Here, both the vector α and the slopes ac are dimensionless quantities. To assess the accuracy of the IM, we compare the predictions of the model against FBA cal- culations for the growth of E. coli on random complex media with a fixed number of non-zero nutrient uptakes (Methods). Because g is defined up to a multiplicative constant, the largest the total uptake, the largest the biomass production. We thus consider complex uptake vectors nor- malized to 1, to mimic physiologic conditions. However, we note that we would obtain the same relative errors for a fixed number of uptakes if we considered non-normalized fluxes. gFBA gFBA−g(φ) Figure 1c shows that despite its simplicity, the idealized model is fairly accurate, with a relative error, ∆ := , ranging from ∼ 0 -- 2% for one nutrient to 24% for 20 uptakes. Note that using Eq. (3) to predict growth lightly overestimates single nutrient contributions to growth, as the corresponding ∆ for growth on one nutrient shows. This effect however is negligible when increasing the number of uptakes above E ≥ 5. It is also apparent that the IM systematically underestimates FBA predictions for media with E ≥ 2 nutrients, which implies that when several nutrients are present, they contribute synergistically to growth. III. RESULTS A. Scaling of second order terms In order to capture nutrient growth synergies, we consider next the second order terms in Eq. (1). Using FBA, we numerically determine βij by setting to zero all entries of the exchange fluxes except φi and φj and computing the difference (cid:0)φi, φj (cid:1) = gFBA(φ(i,j))−αiφi − αjφj, βij (4) where φ(i,j) is the vector φ such that φk = 0 ∀k (cid:54)= i, j (Fig. 2a). Since there is only one output in our system (biomass), the scale of of g is fixed by one of the uptake fluxes (for instance φj) and the dependency on the remaining uptake fluxes can be expressed as dimensionless quantities, which are ratios of uptake fluxes. As a consequence, we expect β to obey a scaling property (Fig. 2b): (cid:18) φi (cid:19) (cid:18) φi (cid:19) 1 φj βij(φi, φj) = βij , 1 ≡ βij φj φj . (5) Remarkably, we find that β displays additional scaling properties. For concreteness, consider the synergy between sugars and fatty acids. We found that the β functions for any sugar -- fatty acid pair (Fig. 2c) collapse on the same curve when the sugar and the fatty acid uptake fluxes φi, φj are rescaled with respect to the effective number of carbons Ci, Cj of the corresponding nutrient (Fig. 2d). One thus has (cid:16) φi (cid:17) φj (cid:48) sug,f acid β = 1 Cj βij (cid:18) Ciφi (cid:19) Cjφj , (6) 5 (7) (cid:48) σiσj β φj = bσj σi tanh bσj σiCjφj so that the introduction of the rescaled β(cid:48) function allows to have a systematic description of growth only given the nutrient -- pair classes, their carbon content and the ratio of their uptake fluxes. For each nutrient -- class pair σ, σ(cid:48) it is therefore possible to define a function β(cid:48) σσ(cid:48) that displays a simple two -- regime behavior (Fig. 2d), in which one of the nutrients becomes the limiting factor in the contribution to growth. Considering again the case of sugars and fatty acids, when the ratio Ciφi/(Cjφj) → 0 the function β(cid:48) sug,f acid grows linearly, while when Ciφi/(Cjφj) (cid:29) 1 it reaches (cid:16) φi a plateau. To capture these two regimes, we propose the generalized phenomenological model: (cid:18) bσiσj Ciφi (cid:19) (cid:17) where bσiσj ≡ lim φi/φj→0 bσj σi ≡ lim φi/φj→∞ β β(cid:48)(φi/φj) φi/φj (cid:48) (φi/φj). , (8) Here σi and σj are the classes of nutrient i, j, respectively, while bσiσj and bσj σi are dimensionless parameters, since they are defined as a flux ratio. These parameters can be interpreted as the limit- ing synergistic contribution to the biomass yield when one of the two nutrients is in excess of the other. In this formulation, knowing the limiting contributions is thus enough to compute the syn- ergistic contribution to growth of any sugar -- fatty acid pair and for any value of the uptake fluxes. For instance, the transition value T (sug, f acid) = bf acid sug/bsug f acid marks the relative sugar -- fatty acid uptake values at which maximal synergy may be attained without waste of nutrients. Figure 3 shows the averaged collapsed curves for all nutrient class pairs we consider. Our calculations indicate that Eq. (7) is a fairly good description for such averaged β(cid:48), although we note that for each nutrient class pair β(cid:48) has different parameters (see Table I and Appendix for a summary of the averaged parameters for each one of these curves). Note that, for nutrients in the same class, it is not necessary to consider all pair permutations. One can, for instance, sort nutrients in a given class σ by their carbon content and evaluate the parameters bσσ only between pairs i, j such that Ci < Cj. This is the approach we follow in evaluating the parameters bσσ(cid:48), which, as a consequence, are not symmetric when σ = σ(cid:48). The phenomenological model in Eq. (7) captures very well the behavior of β(cid:48) for 4 of the 9 cases: (fatty acid, sugar), (fatty acid, fatty acid), (base, sugar) and (base, base) pairs (Figs. 3a, d, b, and g) 2. For the (base, fatty acid) case (Fig. 3 e), we find that the phenomenological model in Eq. (7) does not fully capture the behavior of the averaged β(cid:48) (see Appendix). In such case we still find that β(cid:48) is roughly linear for φ1/φ2 (cid:28) 1 and shows a plateau when φ1/φ2 (cid:29) 1, as predicted by Eq. (7). However, for C1φ1/(C2φ2) (cid:39) 1, the model overpredicts the observed synergy. Despite 2 Note that nutrients in the same class are ordered with their carbon content and pair permutations are not considered. Thus in β(cid:48)( φ1 ), φ1 always corresponds to the nutrient with the smaller number of carbons. This implies that, for φ2 the β(cid:48) within the same class, the average slope and plateau values are not equal (see Table I). We also remind that, in the base-base pair case, we only consider pairs of purines as E. coli cannot catabolize pyrimidines by themselves. 6 this deviation, Eq. (7) is a good trade off between model simplicity and predictive power, since the initial slope of β(cid:48) and the plateau value are well predicted by taking the average of the parameter b over all nutrient pairs. Finally, for all pairs including amino acids (Figs. 3 c,f, h, and i), we find that not all curves collapse into a single one. In particular, we see that when φother/φa.acid (cid:29) 1 ({other: sugar, f acid base, a acid}), the scaling functions reach different plateau values, which always lie either above or below a 10−2 threshold value, respectively. Interestingly, for interclass interactions, any given amino acid consistently reaches a plateau above or below such threshold independent of the other nutrient paired with it. We hence classify amino acids into two groups, L (Low synergy), H (High synergy), according to whether they can attain a synergy below or above the mentioned 10−2 threshold, for interclass synergies. For amino acid-amino acid interactions, we thus divide nutrients into H and L and study intraclass/L-H synergies. This allows us to find two slope and plateau values respectively, each related to the H or L amino acid limiting the interaction in turn. Using a logistic regression model, we find that the set of metabolic pathways in which an amino acid participates determines to which group (H or L) it belongs (see Appendix). By minimizing the Bayesian Information Criterion [23], we see that knowing whether the amino acid participates in the set of six pathways listed in Table II is enough to correctly assign all amino acids except MD- Methionine to either group H or L. Once the corresponding group is known, we can use Eq. (7) to describe β(cid:48) by allowing two plateau values when the nutrient pair involves an amino acid. In this way, we can have close estimates of synergies through the function Eq. (8) for nutrients pairs from all classes, by only knowing their class and the pathways in which they participate. B. Competition for synergistic potentials When a bacterium grows on a complex medium with E > 2 nutrients, Eq. (1) yields a sum over E(E − 1)/2 synergy contributions resulting in an overprediction of the biomass production (see Appendix). The reason for this is that resources are limited by stoichiometry, thus besides the independent nutrient contribution to growth of each uptake φi, resources must be distributed in some way among the E − 1 possible synergies. Two plausible flux allocations are the following: i) an equitative distribution of all {φi} among the synergies (equitative synergy model, ES); ii) a distribution among synergies that yields maximal synergy, which we call optimal synergy model (OS). We find that while the former underpredicts growth rates when increasing the number of uptakes, the latter yields an accurate prediction of FBA growth rates roughly independent of the number of nutrients (Fig. 4 and Appendix). Our results thus suggest that, phenomenologically, one can understand the growth maximization principle observed in microbes as the optimization of nutrient synergies. The OS theory exploits the fact that synergy contributions are limited by the smallest uptake flux Eq. (7), so that only the nutrients in excess can be used in other synergies. In order to maximize the overall synergy, we hypothesize that an optimal allocation of nutrients is adopted to produce the largest pair -- synergies. We thus rank nutrient -- pair synergies and add up to the total synergy each contribution. After each addition, the fluxes of the pair are rescaled such that the limiting one is not considered further, while the nutrient in excess can contribute to other synergies with the fraction of uptake not invested yet (Methods). 7 In a complex growth medium with E non-zero nutrient uptakes, we thus express the OS growth rate as follows: g(φ) = + E(cid:88) P(cid:88) (cid:96)=1 ασ(cid:96)φ(cid:96) (cid:32) (cid:33) , bσσκqrκ bσκσqrκ κ φκCκ  φC (9) bσκσCκqrκ κ φκ tanh (κ,)=1 where the second sum runs over the P = E(E − 1)/2 ranked pairs of nutrients, rκ is the ranking of the nutrient pair synergy (κ, ), and qrκ κ ∈ [0, 1] indicates the fraction of uptake flux φκ yet to be allocated to this contribution. As before, C(cid:96) is the effective number of carbons of nutrient (cid:96) and σ(cid:96) is the nutrient class to which nutrient (cid:96) belongs, and coefficients b have been reported in Table I. The yields ασ(cid:96) can either be directly evaluated for each nutrient, or computed as in equation Eq. (3), with parameters a reported in Fig. 1 b. Note that, when available it is preferable to use the exact α when dealing with less than 4 nutrients, because Eq. (3) slightly overpredicts single nutrient contributions to growth in this case (this effect however vanishes when dealing with E ≥ 5 nutrients). Finally, we compare the biomass production predictions of our OS model Eq. (9) against FBA predictions for E. coli in media with a fixed number of non-zero random nutrient uptakes normal- ized to 1 (Methods). Figure 4a shows the OS model is able to predict with high accuracy the growth rates computed computed by using FBA assuming known uptakes. The average relative error ∆ := over 500 different random growth media with fixed number of uptakes is systematically smaller for OS model predictions than for those of the IM. Notably, the gap between the two models increases with the number of uptakes, due to the more synergistic contributions that are being neglected by the IM model. gFBA−gmodel gFBA Since sugars are the main source of carbons and are quite commonly included in experimental growth media, to reproduce these media we always allow the uptake of one sugar. For more random nutrient setups we find ∆ of the OS to be slightly larger, but still consistently smaller than the IM theory (see Appendix). C. Comparison with experiments After validating our model in silico, we test here how well the OS model predicts actual growth rates in vivo. To do so, we compare our model with experimental measurements of nutrient uptakes and growth for bacterial culture on complex media. Note that obtaining such type of data is 8 generally not straightforward as measurement of multiple uptakes is typically hard. Additionally, to date, standard experiments used to validate FBA generally focus on the simpler case of growth media with a single source of carbon. Nevertheless, a very interesting study on complex media where bacterial growth rate and variation of nutrient concentration are measured was published by Beg et al. [20]. The authors performed there some E. coli batch culture experiments that allowed them to estimate those quantity simultaneously as a function of time. From their published data, we were able to recover the nutrient uptakes corresponding to every measured growth rate (Appendix) and to use such uptakes as inputs in our model. This approach allowed us in turn to compare the predicted growth rate with the experimental one. The results are reported in Fig. 5, where we compare OS model predictions with the experi- mentally measured growth rates. Note that now that physiological uptake and growth values are measured, we can use proper mmol gDW−1h−1 units for the former and h−1 for the latter. When doing so, model Eq. (9) reaches a remarkable accuracy, especially taking into account that i.) the E. coli strain in the experiments differs from the reconstruction at our disposal and ii.) we used the b and a parameters we derived by calibrating the model with FBA, rather than estimating them ad hoc, thus highlighting the broad applicability of our model. The excellent agreement we found between the growth predicted by our model and the actual growth on a complex medium supports that scaling and synergy really are two principles regulating microbial growth in vivo besides their role in modeling metabolism in silico. IV. DISCUSSION: SCOPE AND POTENTIAL LIMITATIONS OF OUR APPROACH We have used FBA predictions under growth optimization as a reliable source of growth rates, that is, as a substitute for growth experiments with real bacteria. Thus, even though our model is ultimately independent of FBA (in that Eq. (9) does not rely in any way on FBA or on any particular metabolic reconstruction), one may argue that our model is susceptible to suffer the shortcomings of FBA. Here we discuss these shortcomings, although the comparison to experimental data in Fig. 5 demonstrates that, whatever limitations FBA may have, our model is able to reproduce experimental growth rates in a variety of realistic conditions. The first issue is the determination of the so-called ATP maintenance flux. This is an additional reaction flux that FBA adds to the set of metabolic reactions and constraints to reproduce the experimental growth rates. Such ATP flux encompasses a series of external factors that affect microbial growth rates, such as the uptake rate of nutrients, oxygen availability, and regulation or temperature. But although ATP maintenance rates obtained for a specific minimal medium have been shown to reproduce accurate results in different growth conditions for certain organisms [24], it cannot be assumed that specific values are valid to make predictions for different growth conditions in general. To overcome this, we proceed as in [19] and first evaluate the ATP needed for the polymerization of biomass components by using the values experimentally determined (which are available in the literature [24, 25]) and then fix the ATP maintenance to this baseline, 9 removing any further ATP maintenance contribution. In any case, it is always possible to rescale our findings a posteriori in the same way ATP maintenance is fitted within the FBA approach. Moreover, Fig. 5 suggests that the effect of the maintenance flux is not very relevant. Another caveat of FBA is that it systematically predicts the simultaneous uptake of different sugars, while it is known that microbes absorb their preferred sugar first [26]. For this reason FBA will regularly over-predict biomass production in presence of multiple sugars [27]. In our approach this is mostly irrelevant because we are concerned with determining growth given the uptakes of nutrients. In any event, to avoid validating our model against unrealistic settings, we focus on complex growth media containing a single sugar (Methods and Appendix). Finally, it has been empirically demonstrated that under certain conditions, unicellular organ- isms do not strictly follow a maximal growth principle [12]. However, it has also been shown that in many occasions the metabolic state predicted by growth maximization is very similar to that of the maximization of other functions [11], so that our formalism could be applicable to these conditions. V. CONCLUSIONS In this work, we present a second order phenomenological model of metabolism that, by relying on a very limited set of parameters, is able to predict the biomass production of E. coli in arbitrary complex growth media within 1% of the actual value for growth in silico and with great accuracy for growth in vivo. Our model shows that nutrients within the same class are effectively catabolized in a similar manner, so that the contribution to growth in the presence of a given nutrient is fully determined by the nutrient's effective carbon content and the class it belongs to. We find that the synergy developed by the uptake of several nutrients increases the catabolic potential of the metabolic network. Such synergy between nutrients pairs depends on the relative abundance of the nutrients and is capped by the less abundant nutrient. Our model shows that, effectively, nutrient contributions to growth can be well approximated by the sum of the independent contribution of each nutrient and a synergy contribution. The syn- ergy contribution depends exclusively on nutrient pair synergies so that uptake fluxes are allocated among pair synergies in order to maximize the synergy contribution with the available resources. In this way, the function maximization principle (usually growth) that determines the metabolic state of a unicellular organism can be effectively understood as the optimization of nutrient syner- gies. METHODS Random flux uptakes generation For each fixed number of uptakes E, we generate a vector φ of uptake fluxes that allows the bacterium 10 to catabolize a combination of fatty acids, amino acids and bases, plus one sugar only. To do so, only one of the entries of φ that do correspond to sugar uptakes is chosen uniformly at random to have a value different from zero. Such value is uniformly drawn at random in the range (0, 1) arb.units. All E − 1 remaining uptakes are uniformly chosen at random among entries of φ that do not correspond to a sugar. Again, the flux value is drawn in the range (0, 1) arb.units. After all the E nonzero entries of φ are drawn, we normalize the uptakes so that the total uptake is always equal to one (see Appendix for results in other complex media). Optimal synergy model Suppose we want to compute the growth of a vector φ of uptake fluxes with E non-zero entries according to the OS model Eq. (9). In order to allocate the uptake of fluxes to maximize synergy we proceed as follows. First, we compute all E(E − 1)/2 synergies β(cid:48) and rank them according to their corresponding contributions to growth from largest to smallest. Starting from the largest, we evaluate which nutrient in the pair (n1, n2) is in excess by comparing the flux ratio Cn1φn1/(Cn2φn2) to the transition value T (n1, n2) = bn2n1/bn1n2 of the corresponding β(cid:48) function. For instance, if Cn1φn1/(Cn2φn2) < T (n1, n2), n2 is in excess. We then store this contribution, set the limiting flux φn1 to zero and reduce φn2 by its distance from the transition value as φn2 → φn2 − Cn1/Cn2φn1T (n1, n2). Note that this implies that φn1 is not used in other synergies. All the other fluxes are kept constant. These updated fluxes are used to re-compute the synergies occupying lower positions in the rank, and the process is repeated for the second largest β(cid:48). In this way synergies at position k in the rank are computed with effective fluxes (φk n2) that take into account both the limitedness of resources and their optimal routing. n1, φk A slightly different version of our approach, where ranking of synergies is computed after each step n → φk+1 φk is not as accurate as the protocol described above (see Appendix and fig. 4). n 11 Figure 1: Idealized metabolism theory. (a) The α parameters introduced in Eq. (2), versus the number of effective carbons for each of the nutrients considered in our study. We consider nutrients in four groups: sugars, fatty acids, bases and amino acids. The α coefficients are a linear function of the effective number of carbons whose slope depends very weakly on the nutrient class, except for bases (see panel b). The dashed lines show linear fits for each class of nutrients, while the black dotted line is a fit considering all of them together. (b) The coefficients ac introduced in Eq. (3). We show the values of ac obtained from the fits shown in panel a). ac varies weakly with nutrient class. (c) Predictions of the idealized metabolism theory, Eq. (2), versus FBA results for a selection of 100 random media with increasing number of possible uptakes (see Methods). Filled red circles correspond to using exact α values, while empty blue squares to Eq. (3). (d) The relative error ∆ = gFBA−gmodel of the IM theory predictions for the two different choices of α averaged over 500 random media, for increasing number of uptakes. ∆ is relatively small in presence of a few nutrients only, but it increases roughly linearly. Note that the error performed when using Eq. (3) in presence of one nutrient only is different from zero, meaning that Eq. (3) does not correctly capture single nutrient contributions to growth. This effect however is negligible increasing the number of nutrients, as the two ∆ curves gFBA overlap. NutrienttypeacSugars1.9×10−2Fattyacids1.5×10−2Aminoacids1.6×10−2Bases0.9×10−2AllNutrients1.9×10−2010203040Effective Carbons0.00.10.20.30.40.50.60.7αSugarsFatty acidsBasesAmino acids05101520No. Uptakes0.000.050.100.150.200.25ΔIndividual NutrientsNutrient classes0.230.470.700.230.470.700.230.470.700.230.470.70^(a)(d)1 Uptake2 Uptakes3 Uptakes≥4 Uptakes(c)gFBA [arb. units]gmodel [arb. units](b) 12 Figure 2: Scaling of nutrient synergy contributions. (a) The function β, Eq. (4), that expresses the gap between the linear model predictions Eq. (2) and the FBA results for the growth rate of E. coli, when there are two nutrient uptakes different from zero. We show here the simultaneous uptake of dodecanoate and butyrate (both fatty acids) as a typical example. β is a growing function of the exchange fluxes of both nutrients. The circles and crosses correspond to the two (example) curves that are shown, once rescaled, in panel (b). (b) Scaling property of β, Eq. (5). We plot the same data points of panel (a): each curve shows β/φ2 as a function of φ1/φ2, for two different fixed values of φ1. Such normalization allows to collapse all points on the same curve. (c) The function Eq. (5) for a set of five sugar-fatty acid pairs, that shows a characteristic linear -- plateau behavior. (d) The rescaling property Eq. (6). We rescale the uptake fluxes of the nutrient pairs shown in panel c with the number of carbons of each nutrient. All the points collapse on the same curve. The dotted line corresponds to the function Eq. (7), where we set bs·fa , bfa·s as the average of the set bs·fa, bfa·s for all the sugar -- fatty acid pairs. ACKNOWLEDGMENTS This work was supported by a James S. McDonnell Foundation Research Award, Spanish Min- isterio de Econom´ıa y Comptetitividad (MINECO) Grant FIS2013-47532-C3, European Union Grant PIRG-GA-2010-277166, European Union Grant PIRG-GA-2010-268342, and European Union FET Grant 317532 (MULTIPLEX) LANA acknowledges the support of NSF award SBE 10-510-410-310-210-1100101β10-310-210-1100101102φ1[mmol gDW-1h-1]10-1100101102103104φ2[mmol gDW-1h-1]10-310-210-1100101φ1/φ210-410-310-2β/φ210-410-310-210-1100101102103φ1/φ210-610-510-410-310-210-1100β/φ2gal/octagal/dcamalttr/octagal/ttdcealyx-L/octa10-410-310-210-1100101102103104φ1C1/(φ2C2)10-610-510-410-310-210-1β'gal/octagal/dcamalttr/octagal/ttdcealyx-L/octa(a)(b)(c)(d) 13 Figure 3: Nutrient synergy contributions. We show the β(cid:48) function, Eq. (6), for pairs of four nutrient classes: sugars, fatty acids, bases and amino acids. Dashed lines correspond to the function in Eq. (7) where the parameters {bκ} are averaged over all pair of nutrients in the corresponding pair of classes. 0624318 Foundation and the W.M. Keck Foundation. 10-710-610-510-410-310-210-1β'10-410-310-210-1100101102103104C1φ1/(C2φ2)10-710-610-510-410-310-210-1β'10-410-310-210-1100101102103104C1φ1/(C2φ2)10-710-610-510-410-310-210-1β'10-410-310-210-1100101102103104C1φ1/(C2φ2)10-710-610-510-410-310-210-1β'Fatty acidsBasesAmino acidsSugarsFatty acidsBasesAmino acids(b)(f)(i)(a)(g)(e)(h)(d)(c) 14 Table I: Average numerical values of the parameters of the phenomenological model in Eq. (7). We show here the average slope (b12) and plateau (b21) values of the β(cid:48) functions for the cross interactions plotted in Fig. 3. For nutrient pairs involving an amino acid we obtain two different plateau values, depending on the metabolic processes in which the amino acid participates (see text). If two amino acids are involved, also an additional slope is needed. When the pair is inverted, for different nutrient classes, the values of the plateau and and slope are also swapped. Note that we order nutrients according to their carbon content and do not consider pair permutations. For this reason, for pairs of the same class (e.g. Fatty acids -- Fatty acids), values b12 and b12 are not equal: b12 captures growth on media where the nutrient with more carbons is in excess, while b12 renders the opposite situation. Sugars Fatty acids 2 1 b12 b21 b12 b21 b12 Fatty acids 2.4 × 10−3 1.2 × 10−2 1.4 × 10−4 3.5 × 10−3 Bases b21 Amino acids b12 b21 Bases 8.8 × 10−4 3.1 × 10−2 1.2 × 10−2 3.4 × 10−2 7.2 × 10−4 1.3 × 10−2 Amino acids 1.6 × 10−3 1.2 × 10−2 3.0 × 10−2 2.9 × 10−3 3.6 × 10−2 3.9 × 10−3 4.1 × 10−2 2.8 × 10−3 1.3 × 10−2 2. × 10−3 5.4 × 10−5 4. × 10−3 3.3 × 10−2 Table II: The metabolic pathways included in the logistic model to predict amino acids groups (H or L). We report in the first column the pathway names, sorted for decreasing Bayesian Information Criterion associated with the model. In the second column we list the number of amino acids participating in each pathway. Metabolic pathway No. a. acids 1. alanine, aspartate and glutamate metabolism 2. valine, leucine and isoleucine degradation 3. phenylalanine, tyrosine and tryptophan biosynthesis 4. sulfur relay system 5. glycine, serine and threonine metabolism 6. arginine and proline metabolism 6 2 3 2 7 7 15 Figure 4: Second order equitative synergy theory. (a) Predictions of the optimized synergy model (OS) Eq. (9), empty blue squares, versus the FBA results, compared with the IM theory Eq. (2), filled red circles, for 100 different random media at increasing number of uptakes (see Methods and Appendix for the details on growth media). Here, we use the exact values of parameter α and the average interclass value of parameters b. (b) The relative error ∆ = gmodel−gFBA vs. the number of uptakes for the IM (filled red circles) and the OS model (empty blue squares), averaged over 500 different random media. The relative error of the IM theory grows almost linearly, while it remains much lower in the gFBA OS model and becomes roughly independent of the number of uptakes for E ≥ 6. 05101520No. Uptakes0.000.050.100.150.200.25ΔIMOS0.230.470.700.230.470.700.230.470.700.230.470.70(b)1 Uptake2 Uptakes3 Uptakes≥4 Uptakes(a)gFBA [arb. units]gmodel [arb. units] 16 Figure 5: Comparison of the OS model, Eq. (9), (y axis) with the experimental growth of Beg et al., [20] (x axis); the dashed diagonal indicates perfect agreement. The uptakes corresponding to each experimental growth rate were computed (Appendix) and used as an input of the OS model to evaluate the predicted growth. The x error bars are one standard error, the y error bars indicate all feasible growths consistent with the uptakes plus/minus their error. We find a fair agreement between our theory and the experimental measurements, supporting that scaling and synergy are two principles regulating also microbial growth in vivo. 00.20.40.60.81gexpt[h-1]00.20.40.60.81gmodel[h-1] 17 Appendix A: The metabolic reconstruction We use the genome scale E. coli metabolic reconstruction iAF1260 [24]. Such reconstruction features 1678 metabolites and 2392 reactions, of which 299 are exchange reactions. The minimal medium is composed by 18 essential nutrients Ca2, cobalt2, Cu2, Zn2, Mn2, cbl1, H2O, Pi, H, K, Cl, Fe2, Fe3, mobd, Na1, Nh4, So4, Mg2 [24]. The fluxes of the reactions that uptake these nutrients are always kept different from zero. In our analysis we assume nutrient uptakes are known. Thus we focus exclusively on the 63 exchange reactions delivering sugars (22 reactions), fatty acids (6 reactions), amino acids (26 reactions), and bases (9 reactions) to the bacterium (see Table III), and keep all other exchanges locked to zero. Appendix B: Flux Balance Analysis Flux Balance Analysis (FBA) is a mathematical tool to predict, under certain assumptions, the fluxes ν and the biomass production gFBA of a metabolic network [9]. Given the stoichiometry S of the network, FBA aims at finding the solution of the metabolic mass balance equation under steady state condition. Denoting by c the vector of metabolic concentration, FBA seeks thus to solve the system of linear equations: c = Sν = 0. (B1) Since in real metabolic networks there are much more reactions than metabolites, the above system is underdetermined and it allows several solutions. From the space of solutions, physiologically relevant points are usually selected by coupling the mass balance problem Eq. (B1) with an opt- mization principle. Quite generally, thus, a FBA problem seeks solutions to Eq. (B1) such that a linear objective function Z of the form (cid:88) Z = rkνk, (B2) k with rk some positive constants, is maximized. The objective function is often related to the biomass production. In our case we focus solely on the maximization of biomass polymerization, so that we have one flux only appearing in the sum Eq. (B2) (which expresses the biomass synthe- sis) and we can assume Z = gFBA. Finally, we note that when essential nutrients are assumed to available in excess, Eq. (B1) specifies a linear problem that is defined up to multiplicative constant: any solution to Eq. (B1) may be rescaled through a constant factor and still be a valid solution. We therefore keep uptakes in arbitrary units when validating our model against FBA. Appendix C: Generation of the growth media We focus only on nutrients that can be uptaken by the organism and produce growth [19]. The growth media we generate therefore only contain sugars, fatty acids, amino acids, and bases. Since Table III: The 63 uptake fluxes considered in our study. We include uptakes delivering sugars (22 reactions), fatty acids (6 reactions), amino acids (26 reactions), and bases (9 reactions) to the bacterium. 18 Sugars 14. Maltose 15. Melibiose 16. Sucrose 17. Trehalose 18. Maltotriose 19. Maltotetraose 20. Maltopentaose 21. 1-4-α-D-glucan 1. L-Arabinose 2. L-Lyxose 3. D-Ribose 4. D-Xylose 5. L-Xylulose 6. D-Allose 7. D-Fructose 8. L-Fucose 9. β-D-Galactose 22. Maltohexaose 10. Galactose 11. D-Mannose 12. L-Rhamnose 13. Lactose Fatty acids 1. Glycine 1. Octanoate 2. D-Alanine 2. Decanoate 3. L-Alanine 3. Dodecanoate 4. Tetradecanoate 4. D-Cysteine 5. Hexadecanoate 5. L-Cysteine 6. Octadecanoate Amino acids 14. D-Methionine 15. L-Methionine 16. Ornithine 17. L-Proline 18. L-Valine 19. L-Arginine 6. D-Serine 7. L-Serine 20. L-Histidine 8. L-Asparagine 21. L-Isoleucine 9. L-Aspartate 22. L-Leucine 10. L-Homoserine 23. L-Lysine 11. L-Threonine 12. L-Glutamine 13. L-Glutamate 24. L-Phenylalanine 25. L-Tyrosine 26. L-Tryptophan Bases 1. Allantoate 2. Cytosine 3. Uracil 4. Adenine 5. Guanine 6. Hypoxanthine 7. Orotate 8. Thymine 9. Xanthine multiple uptake of sugars is not observed [26], we allow for the exchange of one sugar only and randomly allow all other nutrients to be uptaken by the bacterium. Summing up all the exchange fluxes listed in Sec.A, each growth medium can therefore be composed of 42 nutrients at the most (i.e. one sugar and 41 other nutrients), plus the 18 nutrients in the minimal medium. As the minimal medium is always included, just considering the 22 sugars and the 41 remaining nutrients, for each growth medium we hence have a 63 -- dimensional random vector of exchange fluxes φ which, for any fixed number of uptakes E, is generated as follows (see Fig. 6 for a pictorial representation of the growth media): • Only one of the 22 entries delivering sugars is uniformly chosen at random. We randomly fix its value uniformly in the set φsug ∈ (0., 1.) arb.units. • The remaining E − 1 uptakes are uniformly drawn at random among the 41 entries of φ that do not correspond to a sugar. The value of each flux is again uniformly drawn at random in the set (0., 1.) arb.units. • The E nonzero entries of φ are normalized so that(cid:80) i φi = 1 arb.units In all the complex growth media we generate we always include the essential nutrients, which are assumed to be present in excess, i.e. they are uptaken at a rate 1 × 107 arb.units, equivalent to infinite uptake rate in the metabolic reconstruction. 19 Figure 6: Illustration of how random media are generated. Besides the minimal medium, we only consider growth on sugars, fatty acids, amino acids, and bases. Each random medium we generate only contains one sugar (the purple filled arrow), plus a set of other nutrients. The sugar and the remaining nutrients are all uniformly chosen at random. These nutrients and their uptake value form a random vector of exchange fluxes φ. In the figure we sketch as filled arrows all the nutrients included in the random medium and as empty arrows the ones not considered. For any random medium considered, uptakes are normalized so that(cid:80) i φi = 1 arb.units.. Appendix D: Selection of the minimal model for the growth on amino acids When studying nutrient -- class -- wide pairwise interactions involving amino acids, we noticed that the β(cid:48) functions appearing in Fig. 3 tended to acquire two plateau values. We hence divided the amino acids into sets H and L, according to whether their corresponding β(cid:48) plateau value was above or below 10−2, respectively. By doing this, we observed that the pathways that process a given amino acid correlate in some way with its associated β(cid:48) plateau values. Indeed, as we show in Fig. 7, many metabolic pathways feature either amino acids belonging to only one set, or a far exceeding number of amino acids in one of the two sets. We thus opted to predict whether a given amino acid belonged to group H (or L) by exploiting the minimum information on the metabolic processes it participates in. We developed a linear model πi for each amino acid i and used logistic regression to estimate the probability Pi(i ∈ Hπi) for metabolite i to belong to group H given model πi. Considering a set M of n metabolic E. coli iAF1260External mediumSugarsFatty acidsAmino acidsBasesφ 20 Figure 7: Number of amino acids in sets H and L for each metabolic pathway. We see that the amount of amino acids in each set is uneven in the majority of pathways, with most of them only featuring amino acids in the L set. We opted to exploit this characteristic to predict to which set each amino acid belongs to and automatically assign it a β(cid:48) pathways, we assumed plateau value. πi ≡ ξ0 + n(cid:88) j=1 1 ξjX j i (D1) Pi(i ∈ Hπi) = , 1 + exp πi where the sum runs over the n pathways in M. In Eq. (D1) X j i is a binary variable taking value 1 if amino acid i participates to pathway j and 0 otherwise. All coefficients {ξj}n j=1 have real values. For each set M we estimate {ξj}n i=1, Pi. The coefficient ξ0 is related to the probability that an amino acid i belongs to H while not participating to any pathway in πi. As we aim to gain the maximum predictive power by exploiting the minimum information, we opted to seek for the smallest set M that yields the largest rate of correct guesses, j=1 by miximizing the likelihood L =(cid:81) abc transportersaminoacyl-trna biosynthesiscysteine and methionine metabolismd-glutamine and d-glutamate metabolismc5-branched dibasic acid metabolismbeta-alanine metabolismlysine biosynthesisalanine, aspartate and glutamate metabolismsphingolipid metabolismpurine metabolismtyrosine metabolismglutathione metabolismcyanoamino acid metabolismarginine and proline metabolismpyrimidine metabolismporphyrin and chlorophyll metabolismnicotinate and nicotinamide metabolismsulfur metabolismthiamine metabolismbutanoate metabolismmicrobial metabolism in diverse environmentsd-alanine metabolismbacterial chemotaxisselenocompound metabolismpantothenate and coa biosynthesisphenylalanine metabolismmethane metabolismglycine, serine and threonine metabolismhistidine metabolismnitrogen metabolismpeptidoglycan biosynthesisnovobiocin biosynthesissulfur relay systemlysine degradationtwo-component systemvitamin b6 metabolismtaurine and hypotaurine metabolismphenylalanine, tyrosine and tryptophan biosynthesisfolate biosynthesisvaline, leucine and isoleucine biosynthesisglycerophospholipid metabolismtryptophan metabolismbiosynthesis of secondary metabolites22446688Number of Amino acidsL SetH Set 21 that is, which returns Pi larger than 0.5 for metabolites actually belonging to H in the majority of cases. The minimum set may be found by minimizing the Bayesian information criterion (BIC) [23] , viz: BIC = (n + 1) log N − 2 log L, (D2) where n ≡ (cid:107)M(cid:107) is the size of the set M (i.e. the number of included pathways), N is the number of amino acids and L is the likelihood that the observed H, L sets are generated by models {πi}N i=1. To seek for the minimal M, we started out with zero pathways and then used an iterative greedy approach that at each step added the pathway that yielded the minimum BIC, that is, that maximized the likelihood L. The result of this iterative approach is shown in Fig. 8: the first point features one metabolic pathway and renders a BIC close to 30. Adding parameters (i.e. adding metabolic pathways) lowers the BIC up to n = 6 where there is no more significative gain in predictive power and adding more pathways only overfits the model, so that the BIC starts to grow. The whole analysis was performed using R (version 2.15.3 [28]). Once we knew the profile of the BIC, we retained the set M that minimized it. Such set is the best trade off between the likelihood L (i.e. the predictive power) and the number of pathways included in the model. The six pathways included in the final M yielded a BIC = 27.3 and are listed in Table IV, where we also report the BIC returned by all models featuring n ≤ 6 pathways and the number of amino acids participating in each pathway included. In Fig. 9, we show the probabilities Pi(i ∈ Lπi) as a function of the number of pathways n in the model πi. In our analysis we fix a threshold of 0.5 and assume metabolite i belongs to H if Pi > 0.5 and i ∈ S otherwise. The green shaded area in Fig. 9 indicates the region where we expect Pi to lie: for the vast majority of the amino acids only a few parameters in the πi are sufficient to classify all amino acids into sets L or H. For the case n = 6 pathways, which minimizes the BIC, we see that there is only one amino acid which is not correctly classified, namely D-Methionine (met D). All the rest of the amino acids are correctly assigned to either L or H by only inspecting whether they participate in the metabolic pathways listed in Table IV. Since knowing whether a given amino acid participates to these six pathways is sufficient to know where its associated β(cid:48) plateau will lie, we decided to model the β(cid:48) functions through their phenomenological form Eq. (8) and assign two possible values to parameters b, which are evalu- ated by averaging β(cid:48) corresponding to amino acids in the sets H and L separately. Appendix E: Optimal synergy in the second order model As shown in Sec. II, the IM model systematically underpredicts growth rates in presence of multiple nutrients. As a result we have to include a synergy term in our model. We do so by introducing the β(cid:48) functions. However, we find that an equal contribution of all synergisitc terms overpredicts the growth rate in complex media (see fig 10). This is because resources are limited and not all nutrient pairs can develop such maximal synergy. We therefore call this a naive eq- 22 Figure 8: The Bayesian Information Criterion as a function of the number of pathways n. Starting with zero pathways, we iteratively incorporated into the model Eq. (D1) the metabolic pathway that yielded the minimum BIC. This allows to gain predictive power and to lower the BIC up to n = 6 pathways (black arrow). Inclusion of further information does not enhance the predictive ability and only overfits the model. Table IV: The six pathways included in the model π that minimizes the Bayesian Information criterion. We report in each row the name of the pathway, the number of amino acids participating in it, and the BIC value of the model containing all pathways up to the row, so that the last line has the minimum BIC value. BIC Metabolic pathway no. a. acids 34.0 alanine, aspartate and glutamate metabolism 33.3 valine, leucine and isoleucine degradation 31.3 phenylalanine, tyrosine and tryptophan biosynthesis 29.8 sulfur relay system 29.1 glycine, serine and threonine metabolism 27.3 arginine and proline metabolism 6 2 3 2 7 7 uitative synergy (NES) model, that assuming maximal synergy among all nutrients describes an unrealistic scenario. In order to limit the overall synergy, we tested the equitative synergy (ES) theory, where re- sources are equally distributed across the nutrient pairs. We created complex growth media as 0102030Number of pathways, n406080BIC 23 Figure 9: The probabilities Pi(i ∈ Hπi) of each amino acid i varying the number of pathways n included in the model πi. The shaded green area highlights the expected region where Pi should lie, i.e. Pi ∈ [0, 0.5] and Pi ∈ (0.5, 1] for amino acids in sets L and H respectively. For the majority of them, the inclusion of only a few pathways in πi is enough to predict the correct set. When n = 6, that is, when the BIC is minimum, we correctly capture the behavior of all amino acids except for D-Methionine (met D). 00.51Pi00.51Pi00.51Pi00.51Pi00.51Pi01530No. Pathways00.51Pi01530No. Pathways01530No. Pathways00.51Pi01530No. Pathwaysval_Ltrp_Lleu_Larg_Lphe_Lthr_Lcys_Dpro_Lasn_Lcys_Lorntyr_Lasp_Lgln_Lala_Dser_Dglu_Lser_Llys_Lglymet_Lile_Lala_Lhis_Lmet_Dhom_L 24 explained in Sec. C, with each medium κ consisting of Eκ nutrients and thus Pκ = Eκ(Eκ − 1)/2 i was equally invested in possible pairs. We then assumed that, for each nutrient i, the uptake φκ the Eκ − 1 synergies such nutrient can develop. Therefore, we computed the ES model growth on medium κ by correcting the IM theory with the β(cid:48) contributions Eq. (7) as: bσj σiCjφκ j bσiσj Ciφκ i φκ i Cibσiσj tanh (cid:88) gκ ES = gκ IM + (E1) 1 Eκ − 1 i<j . Here gκ Pκ possible nutrient pairs. Hence, with factor 1/(Eκ − 1), we equally spread φκ synergies. IM is the IM theory growth, Eq. (2), σi is the class of nutrient i, while the sum runs on the i across the Eκ − 1 The resulting model shows an improvement respect to the IM theory, although the gain de- creases when the number of uptakes grows. The decrease in accuracy for increasing E of both the NES and the ES model suggests that the uptake of resources is distributed in some optimal way. Since in the FBA approach metabolism is aimed at growth optimization, we hypothesized that uptakes are organized in such way to maxi- mize the nutrient synergistic contributions to growth. Specifically, such optimality must be reached by considering that nutrient uptakes that are invested to attain a certain synergy may not contribute to another synergy. In Fig. 3, one clearly realizes how this can be taken into account. Indeed, the β(cid:48)(Cn1φn1/(Cn2φn2)) functions shown in Fig. 3 typically have a growing regime followed by a plateau. The appearance of the plateau means that the synergy is not affected by a variation of the uptake of nutrient n1, i.e. nutrient n1 is in excess with respect to nutrient n2. Conversely, in the growing region, the situation is reverted and nutrient n2 is in excess. The point T (n1, n2) = b21/b12 marks the transition from one regime to the other. Thus, if Cn1φn1 < Cn2φn2b21/b12, nutrient n2 is in excess: in such case, n1 has been completely invested and it cannot be used in other synergies, while n2 can only contribute further with an effective flux Cn2φ(cid:48) n2 = Cn2φn2 − Cn1φn1T (n1, n2) 3, that is, with the surplus of its uptake. We hence devised the following method to achieve optimality in the case of limited resources on complex growth media: 1. For each pair of nutrients i, j and corresponding uptake fluxes φi, φj compute the second order correction ∆gij to the IM growth: ∆gij = Cjφjbσj σi tanh bσiσj Ciφi bσj σiCjφj , (E2) where σi and Ci are the class and the carbon content of nutrient i, respectively. 2. Rank all ∆gij from largest to smallest. The first in such rank will be the best contribution to accomplish optimal growth. 3. Add to the IM growth prediction the first correction in the rank. 3 Cn1φ(cid:48)n1 = Cn1 φn1 − Cn2φn2 /T (n1, n2), φn2 = 0 respectively if nutrient n1 is in excess 25 Figure 10: Second order model predictions. (a) Prediction of model bacterial growth against FBA results, for four models (see text): IM, NES, ES, OS. The idealized metabolism (IM, red circles) captures reasonably well FBA growth predictions. Including maximal synergy for all the nutrient pairs with a naive equitative synergy theory (NES, purple up triangles) largely overestimates the FBA growth. Considering a uniform uptake for all nutrient pairs with the equitative theory (ES, green diamonds) improves the IM results. When the number of uptakes is (cid:29) 1, all these models produce worse results than the Optimized Synergy model (OS, blue squares). (b) The relative error ∆ of the different models as a function of the FBA growth gFBA. The baseline is the first order IM theory (red circles), with a relative error that increases roughly linearly with the number of uptakes. The NES model (purple up triangles) is clearly unrealistic, with a relative error that increases very fast. The ES model (green diamonds), conversely, improves the IM results, although its ∆ still increases with the number of uptakes. The OS model error (blue squares) remains very low and depends very weakly on the number of uptakes, suggesting optimal allocation of synergies is a robust explanation for maximal growth. 05101520No. Uptakes0.000.050.100.150.20ΔIMNESESOS0.230.470.700.230.470.700.230.470.700.230.470.70(b)1 Uptake2 Uptakes3 Uptakes≥4 Uptakes(a)gFBA [arb. units]gmodel [arb. units] 26 4. Reduce fluxes φi and φj, so to take into account that some uptake of nutrients i and j has been invested into their synergy: (a) For the nutrient in excess, say j, set φj → φj − Ci/Cjφibσj σi/bσiσj. (b) Set φi → 0, as uptake of i has all been used to develop synergy ∆gij. 5. Remove from the rank all synergies involving nutrient i, as its effective uptake is now zero. 6. Re-compute the synergies {∆gkj} with the new uptake flux φj. 7. Optimal synergy (OS) model: go to step 3. The process is iterated until no uptake flux can be diminished further. The above strategy to pinpoint optimal allocation of resources is really effective. The OS model gives very accuarate results even for a large number of uptakes and we thus opted for it. Note that the results presented are derived assuming that a sugar is always present in the medium. One can generalize and also work with sugar-free complex growth media. Because β(cid:48)(x) functions for (fatty acid, base), (fatty acid, amino acid), and (amino acid, amino acid) inter- actions are not perfectly captured by Eq. (7) when x (cid:39) 1, this scenario is better captured allowing for two different slopes of the beta functions: results for the OS model are slightly less accurate than in presence of sugars, but still far better than the IM, as shown in Fig. 11. Figure 11: (a) Predictions of the OS model (blue open squares) vs the IM model (red filled circles), for complex media that may not include sugars. To better capture non -- sugar synergies we allow here 2 different slopes to the β functions (b) The relative error ∆ of the OS model (blue empty squares) and the IM model (red filled circles). Also when sugar are not always uptaken the OS model has a consistently smaller relative error than the IM model. 05101520No. Uptakes0.000.050.100.150.200.25ΔIMOS0.240.470.710.240.470.710.240.470.710.240.470.71(b)1 Uptake2 Uptakes3 Uptakes≥4 Uptakes(a)gFBA [arb. units]gmodel [arb. units] 27 Appendix F: Comparison with the experiments Beg et al. [20] published a few years ago a study that proves to be an excellent means to contrast our model against experimental results. In their work, the authors measured at high frequency the growth rate of a batch culture of E. coli and the corresponding variation of nutrient concentration in the medium, simultaneously. Additionally, they included in their paper measurements of the culture optical density and other quantities of interest. All the relevant measurements for our analysis are reported in Ref. [20] Fig. 2, panels a and b: in the following, we explain how to integrate such data in our approach. The first step to make the results of Beg et al. useful in our framework is to calculate, for each nutrient i, the uptakes φi given the time evolution of nutrient concentration ci(t) reported in Fig. 2b of Ref. [20]. For each nutrient i, the uptake φi is related to the time derivative of the nutrient concentration ci as: φi(t) = VW ci(t), (F1) 1 D(t)mi where mi is the molar mass of nutrient i, D(t) the microbial dried mass at time t and VW is the working volume, which is provided by the authors in the supporting material of Ref. [20] (note indeed that concentration are provided per unit volume in [20]). This relation properly yields uptakes in mmol gDW−1 h−1, the units commonly applied in metabolic reconstructions and that we use in our model. From Eq. (F1), we see that, to compute φi(t), first the derivatives ci must be evaluated from the provided curves ci(t), for each nutrient i. This is straightforward and can be accomplished with, e.g., centered differences. For each value ci(t) we also compute the error σ ci(t) evaluating the maximum and minimum slopes compatible with the given error bars of ci(t), also reported in Fig. 2b of Ref. [20]. The second quantity to evaluate in order to calculate the uptakes is the dried weight D(t). We assume it to be proportional to the optical density O(t), which is given in Fig. 2a of Ref. [20]. Knowing the initial optical density O(0) and dried weight D(0) (which is specified to be 6.75 × 10−3 g), we are hence able to compute the whole D(t) curve, with its own error σD(t) (evaluated from the known error on the optical density). After the above step, we are able to compute the uptakes φi(t) and their associated errors σφi(t) (propagating σ ci(t) and σD(t)), for each nutrient i and time t. Note that we do not allow negative uptakes (corresponding to nutrient release, really) and we discard noisy fluctuations of ci(t) allowing for unexpected multiple nutrient uptakes at t ≤ 3.5h. Consequently, φi(t) = 0 with zero uncertainty for all nutrients except glucose when t ≤ 3.5h. The resulting uptakes are plotted in Fig. 12a. Knowing all uptakes for each time t, we finally compute the growth gOS(t) predicted by the OS model by using Eq. (9). We also derive an associated error σgOS(t) by evaluating the growth rates yielded by the minimum φmin(t) = {φi(t) − σφi(t); i ∈ nutrients} and maximum φmax(t) = {φi(t) + σφi(t); i ∈ nutrients} possible uptake vectors, respectively. Therefore, in turn, σgOS(t) = (cid:0)φmax(t)(cid:1) gOS (cid:0)φmin(t)(cid:1). 28 − gOS Albeit the experimental growth rate is partially provided in Fig. 2a of Ref. [20], we opt to calculate the experimental growth rate gexpt(t) resulting from our estimate of the experimental dried weight curve D(t). The rationale is to have a gexpt(t) consistent with the D(t) values used to compute the uptakes. Note indeed that in Fig. 2a of Ref. [20] the entire time series of the experimental growth rate is not available (i.e. time window t = 0 to t = 1.5h is missing), so we cannot proceed the other way around and estimate D(t) integrating back the growth rate. Hence, we evaluate gexpt(t) from the differential equation: gexpt(t) = D(t) D(t) , (F2) that fixes the evolution of the dried weight in exponential growth condition. Again we estimate D(t) from D(t) with centered differences and its error σ D(t) analogously to what done for σ c(t). Finally, we compute the error σgexpt(t) for gexpt(t) by propagating σ D(t) and σD(t). The growth rates gexpt(t) we find are entirely consistent with the ones originally published in Fig. 2a of Ref. [20], as shown in Fig. 12b. However, as said, such gexpt(t) values are more coherent with the dried weight we used in Eq. (F1) to compute the uptakes, so these are the ones we plot in Fig. 5. Having computed gOS(t) and gexpt(t), we finally compare them in Fig. 5, finding an excellent agreement. To obtain these accurate results, we use Eq. (3) to estimate the value oif each α. In Fig. 13 we show how results change when using the exact α values instead: the predictions are only slightly better. This finding is remarkable, because to use Eq. (3) we only need to use the slopes ac (Fig. 1b) and the carbon content of each nutrient, rather than the actual yield. The ac values hold for all nutrients in a given class, while the carbon content of nutrients is generally known, so that Eq. (3) can be readily applied to diverse situations without having to reevaluate single nutrient contributions to growth. Note that in these two validations against experimental results we only focus on the truly expo- nential growth phase, i.e. where tgexpt(t) (cid:38) 1, which is the shaded region in Fig. 12. A final remark on the fact that the experimental growth medium contains lactate and glycerol, which do not belong to nutrient classes we discuss presently. Again, one can proceed as we outline in Secs. II and III A to evaluate parameters a and b for the classes corresponding to these nutrients. For organic acids, the class lactate belongs to, we find aorg ac = 1.5× 10−2, while b parameters for all cross interactions are reported in Table V. For glycerol, we opt instead to use the same a and b parameters we derived for fatty acids, which do yield accurate results already. [1] L. B. Ray, Science 330, 1337 (2010). [2] C. S. Henry, M. DeJongh, A. A. Best, P. M. Frybarger, B. Linsay, and R. L. Stevens, Nat Biotechnol 28, 977 (2010). [3] N. Christian, P. May, S. Kempa, T. Handorf, and O. Ebenhoh, Molecular bioSystems 5, 1889 (2009). 29 Figure 12: (a) The experimental uptakes φ computed via Eq. (F1), for the five nutrients considered in Fig. 2b of Beg et al. [20]. Glucose is almost totally consumed first, the rest of nutrients is consumed for t > 3.5 h. Note that the dried weight, which normalizes the plotted values, steadily grows in time. The grey shaded area is the purely exponential growth time window (tgexpt(t) (cid:38) 1), where we pick the points plotted in Fig. 5. (b) Comparison of the growth rate gexpt(t) calculated via Eq. (F2) (Calc., red circles) and the values directly published in Fig. 2a of Ref. [20] (Publ., blue squares). The two quantities are fully consistent, all points but one being within one standard error. We use the values corresponding to the red circles to validate our model in Fig. 5, as they are also related to the dried weight employed to compute the nutrient uptakes. The shaded area once again denotes the pure exponential growth region. [4] J. D. Orth and B. Palsson, BMC systems biology 6, 30 (2012). [5] M. A. Oberhardt, B. O. Palsson, and J. A. Papin, Molecular systems biology 5, 320 (2009). [6] A. Varma and B. Palsson, Nature Biotechnology 12, 994 (1994). [7] D. Segr`e, D. Vitkup, and G. M. Church, Proc. Natl. Acad. Sci. USA 99, 15112 (2002). [8] K. J. Kauffman, P. Prakash, and J. S. Edwards, Current Opinion in Biotechnology 14, 491 (2003). [9] J. Orth, I. Thiele, and B. Palsson, Nature biotechnology 28, 245 (2010). [10] R. L. Chang, K. Andrews, D. Kim, Z. Li, A. Godzik, and B. O. Palsson, Science 340, 1220 (2013). [11] R. Schuetz, N. Zamboni, M. Zampieri, M. Heinemann, and U. Sauer, Science (New York, N.Y.) 336, 601 (2012). [12] S. Bordel, Scientific reports 3, 3017 (2013). [13] F. G. Bader, Biotechnol Bioeng 20, 183 (1978). [14] T. Egli, U. Lendemann, and M. Snozzi, Antonie Van Leeuwenhoek 63, 289 (1993). [15] K. Kov´arov´a-Kovar and T. Egli, Microbiol Mol Biol Rev 62, 646 (1998). [16] K. Toda, J Gen Appl Microbiol 49, 219 (2003). [17] M. Zinn, B. Witholt, and T. Egli, J Biotechnol 113, 263 (2004). [18] V. M. Boer, C. A. Crutchfield, P. H. Bradley, D. Botstein, and J. D. Rabinowitz, Molecular biology of the cell 21, 198 (2009). 02468time[h]0102030 φ[gDW-1h-1]glucoselactatemaltosegalactoseglycerol02468time[h]00.20.40.60.81gexp[h-1]Calc.Publ.(a)(b) 30 Figure 13: Model prediction of experimental growthg rates. We compare here the accuracy of model Eq. (9) at predicting experimental bacterial growth rates when using Eq. (3) to estimate the α parameters (red circles) and by using the exact values of α (blue squares), which are evaluated by estimating the nutrients yield. Eq. (3) performs fairly well, its predictions being only slightly worse than the ones obtained with the exact αs. This is remarkable, as it implies that, when dealing with physiological values, one can accurately predict growth rates by only knowing the slope ac of each nutrient class and the carbon content of each nutrient, respectively, rather than the exact yield. Table V: The OS model b parameters for synergies with organic acids (org ac in the Table). Interactions with amino acids again allows for two different plateau values of the β(cid:48) function. Nutrients are always sorted for increasing carbons: organic acids intra -- class interaction does not consider pair permutations and yields thus two different b values: borg ac other corresponds to growth on a medium where large carbon content organic acids are in excess, while bother org ac captures the opposite situation. other borg ac other bother org ac Sugars Fatty acids Bases 3.0 × 10−3 3.4 × 10−3 Organic acids 2.7 × 10−3 2.4 × 10−3 Amino acids 3.0 × 10−3 3.0 × 10−3 1.8 × 10−2 1.7 × 10−3 1.1 × 10−2 4.0 × 10−3 2.4 × 10−2 00.20.40.60.81gexpt[h-1]00.20.40.60.81gmodel[h-1]α = Eq. (3)α Exact 31 [19] S. M. D. Seaver, M. Sales-Pardo, R. Guimer`a, and L. A. N. Amaral, PLoS Comput Biol 8, e1002762 (2012). [20] Q. K. Beg, A. Vazquez, J. Ernst, M. A. de Menezes, Z. Bar-Joseph, A.-L. Barab´asi, and Z. N. Oltvai, Proc Natl Acad Sci U S A 104, 12663 (2007). [21] M. Durot, P.-Y. Bourguignon, and V. Schachter, FEMS Microbiology Reviews 33, 164 (2009). [22] E. Almaas, B. Kov´acs, T. Vicsek, Z. N. Oltvai, and A.-L. Barab´asi, Nature 427, 839 (2004). [23] G. Schwarz, Ann. Stat. 6, 461 (1978). [24] A. M. Feist, C. S. Henry, J. L. Reed, M. Krummenacker, A. R. Joyce, P. D. Karp, L. J. Broadbelt, V. Hatzimanikatis, and B. Ø. Palsson, Mol. Syst. Biol. 3, 121 (2007). [25] F. Neidhardt, J. Ingraham, and M. Schaechter, Physiology of the bacterial cell: a molecular approach (Sinauer Associates, Sunderland, MA, 1990). [26] J. Monod, Endocrinology 78, 412 (1966). [27] H. Dong, L. Nilsson, and C. G. Kurland, J. Mol. Biol. 260, 649 (1996). [28] R Development Core Team, R: A Language and Environment for Statistical Computing, R Foundation for Statistical Computing, Vienna, Austria (2005).
1901.01281
1
1901
2019-01-04T19:11:58
Calibrating evanescent-wave penetration depths for biological TIRF microscopy
[ "physics.bio-ph", "q-bio.QM" ]
Roughly half of a cells proteins are located at or near the plasma membrane. In this restricted space the cell senses its environment, signals to its neighbors and ex-changes cargo through exo- and endocytotic mechanisms. Ligands bind to receptors, ions flow across channel pores, and transmitters and metabolites are transported against con-centration gradients. Receptors, ion channels, pumps and transporters are the molecular substrates of these biological processes and they constitute important targets for drug discovery. Total internal reflection fluorescence microscopy suppresses background from cell deeper layers and provides contrast for selectively imaging dynamic processes near the basal membrane of live-cells. The optical sectioning of total internal reflection fluorescence is based on the excitation confinement of the evanescent wave generated at the glass-cell interface. How deep the excitation light actually penetrates the sample is difficult to know, making the quantitative interpretation of total internal reflection fluorescence data problematic. Nevertheless, many applications like super-resolution microscopy, colocalization, fluorescence recovery after photobleaching, near-membrane fluorescence recovery after photobleaching, uncaging or photo-activation-switching, as well as single-particle tracking require the quantitative interpretation of evanescent-wave excited images. Here, we review existing techniques for characterizing evanescent fields and we provide a roadmap for comparing total internal reflection fluorescence data across images, experiments, and laboratories.
physics.bio-ph
physics
Oheim et al. (2019) TIRF calibration Calibrating evanescent-wave penetration depths for biological TIRF microscopy Short title: TIRF image quantification Martin Oheim, *,†,‡,1 *, Adi Salomon, ¶,2 Adam Weissman, ¶ Maia Brunstein, *,†,‡,§ and Ute Becherer£ * SPPIN -- Saints Pères Paris Institute for the Neurosciences, F-75006 Paris, France; † CNRS, UMR 8118, Brain Physiology Laboratory, 45 rue des Saints Pères, Paris, F-75006 France; ‡ Fédération de Recherche en Neurosciences FR3636, Faculté de Sciences Fondamentales et Biomédicales, Université Paris Descartes, PRES Sorbonne Paris Cité, F-75006 Paris, France; ¶Department of Chemistry, Institute of Nanotechnology and Advanced Materials (BINA), Bar-Ilan University, Ramat-Gan, 5290002, Israel; §Chaire d'Excellence Junior, Université Sorbonne Paris Cité, Paris, F-75006 France; £Saarland University, Department of Physiology, CIPMM, Building 48, D-66421 Homburg/Saar, Germany; * Address all correspondence to +33 1 4286 4221 (Lab), -4222 (Office) +33 1 4286 3830 [email protected] Dr Martin Oheim SPPIN -- Saints Pères Paris Institute for the Neurosciences 45 rue des Saints Pères F-75006 Paris Phone: Fax: E-mails: 1) MO is a Joseph Meyerhof invited professor with the Department of Biomolecular Sciences, The Weizmann Institute for Science, Rehovot, Israel. 2) AS was an invited professor with the Faculty of Fundamental and Biomedical Sciences, Paris Descartes University, Paris, France during the academic year 2017-18. arXiv - presubmission 1 Oheim et al. (2019) TIRF calibration ABSTRACT. Roughly half of a cell's proteins are located at or near the plasma membrane. In this restricted space, the cell senses its environment, signals to its neighbors and ex- changes cargo through exo- and endocytotic mechanisms. Ligands bind to receptors, ions flow across channel pores, and transmitters and metabolites are transported against con- centration gradients. Receptors, ion channels, pumps and transporters are the molecular substrates of these biological processes and they constitute important targets for drug discovery. Total internal reflection fluorescence (TIRF) microscopy suppresses background from cell deeper layers and provides contrast for selectively imaging dynamic processes near the basal membrane of live-cells. The optical sectioning of TIRF is based on the exci- tation confinement of the evanescent wave generated at the glass/cell interface. How deep the excitation light actually penetrates the sample is difficult to know, making the quantitative interpretation of TIRF data problematic. Nevertheless, many applications like super-resolution microscopy, co-localization, FRET, near-membrane fluorescence recovery after photobleaching, uncaging or photo-activation/switching, as well as single-particle tracking require the quantitative interpretation of EW-excited images. Here, we review existing techniques for characterizing evanescent fields and we provide a roadmap for comparing TIRF data across images, experiments, and laboratories. (193 words) KEYWORDS: nanometer, near-field, fluorescence, excitation confinement, optical sectioning, near-membrane. arXiv - presubmission 2 Oheim et al. (2019) TIRF calibration Quantifying total internal reflection fluorescence The purpose of TIRF is to selectively illuminate fluorophores that are right near a surface and not illuminate fluorophores that are further into the volume, above the surface. Major uses of TIRF include single-molecule detection (SMD) -- e.g., for the localization-based super-resolution microscopies -- as well as studies of cell-substrate contact region of living cells grown in culture. Typical biological applications are the investigation of cell adhesion sites, of the dynamics of membrane receptors, single-vesicle exo- and endocytosis or ER/plasma-membrane contact sites. The confinement of excitation light to a thin, near-inter- face layer results in background reduction and contrast enhancement. Many TIRF applications require knowing the depth of the illuminated layer, e.g., for quantifying the motion of molecules or cellular organelles near the substrate. In principle, that depth can be calculated from the local refractive index (RI) of the medium (n1) and the incidence angle of the illumination (θ) at a known wavelength (λ). But in practice, the actual depth is much less certain. The incident light is often not perfectly collimated but in a range of angles, which sets up a continuous range of penetration depths. Also, the optics used to guide the incident beam upon the sample itself scatters light, some of which enters the sample as propagating light, which does not decay exponentially with distance from the TIR surface. Another source of uncertainty in TIR depth is that the sample's refractive index is not known precisely, and it may contain non-homogeneities, producing scattering. In this perspective article, we briefly recall the fundamentals of evanescent waves (EWs) prior to discussing the concepts, methods and difficulties of calibrating TIRF in- tensities. We suggest a protocol for TIRF image quantification to better control, compare and share results. Other aspects of TIRF and TIRF microscopy have been covered in a number of excellent reviews and tutorials [1; 2; 3; 4; 5; 6; 7; 8]. How to 'see' EWs? Light impinging at an interface between two mediaa having, respectively, RIs n1 and n3 with n3 > n1 is partially reflected and partially transmitted. Snell's and Fresnel's laws govern, respectively, the angles and intensities of the reflected and refracted beams. A discontinuity is observed at a very oblique angle, above the critical angle θc = asin(n1/n3) at which the reflected intensity becomes equivalent to that of the incident beam. Yet, there is light in the rarer medium (n1), because energy and momentum conservation at the boundary prescribe a an intermediate layer having a RI n2 and thickness d2 is for the moment neglected. arXiv - presubmission 3 Oheim et al. (2019) TIRF calibration the existence of a near field, skimming the surface. This evanescent (vanishing) wave propagates on the surface (as does the refracted beam for θ = θc) but its intensity is decaying exponentially perpendicular to the surface. Typical decay lengths are a fraction of the wavelength and can be as small as the λ/5, depending on the angle of incidence, θ, Fig. 1A. EWs are non-visible in the far field, but scattering or absorption couple out energy from the near field, measurable either as an intensity loss in the reflected beam ('attenuated' or 'frustrated' TIR), or as fluorescence excitation in a thin boundary layer above the reflecting interface (TIRF). EW scattering at RI-discontinuities at or near the interface results in far-field light, too. In either case, the total intensity depends on how deep the EW reaches into the rare medium (n1). Another way to 'see' EWs is by their coupling to collective electron oscillations ('plasmons') in a thin metal film (n2, d2) deposited on the glass. Momentum matching is only possible due to the foreshortening of the EW wave vector (wave-front squeezing) and it leads to a sharp decrease in the reflected intensity at the surface plasmon resonance (SPR) angle, θSPR > θc. Changes in n3, e.g., by the binding of molecules to the surface, will shift θSPR. This angle shift is the basis for SPR-based sensing, spectroscopy, and SPR microscopy of cell/substrate contact sites [9; 10; 11]. (Figure 1 Light confinement by an evanescent wave close to here) The theoretical framework For a beam impinging from the left, the EW propagates along +x. Its intensity decays in axial (z-) direction, with -- in theory -- an exponential dependence, I(z;θ) = I0(θ) exp[-z/δ(θ)]. The distance over which the intensity drops to 1/e (37%) of its value at the interface (z = 0) is called the 'penetration depth', (Eq. 1) 𝛿𝜃 = !! !!!!"#!!!!!!= ! ! !!!! !"#!!!!"#!!! It depends on the excitation wavelength λ, the polar beam angle θ and the RIs, Fig. 1A. We here omitted possible intermediate layers (nj, dj) [12; 13]. The penetration depth, δ, does neither depend on the polarization nor on the azimuthal angle φ of the incoming beam. On a plot of δ(θ) vs. θ, Fig. 1B, we recognize an asymptotic behavior, both for θ → θc, for which arXiv - presubmission 4 Oheim et al. (2019) TIRF calibration approaches 𝜆 4𝜋𝑛! 1−sin!𝜃! δ diverges and becomes infinite (which is intuitive, because of the emergence of the transmitted, refracted beam for θ <θc), and for very large θ → π/2, for which δ(θ) , Fig. 1C. For a given λ, higher-index substrates (n3) or more grazing angles θ result in a better optical sectioning. Many unknowns affect the true penetration depth As there is a smooth intensity roll-off, the actual penetration depth is strictly defined only in relation to a certain signal-to-background or signal-to-noise level. For a mono-exponential, I(z) decays to <5% of I(0) over an axial distance of two penetration depths, z = 2δ. With some arbitrariness, one could thus take 2δ as the effective probe depth, because 5% is of the order of typical noise levels. Unfortunately, the question of how far excitation light reaches into the cell is more complicated, and it has been a matter of passionate debate (see, e.g., http://lists.umn.edu/cgi-bin/wa?A2=ind1011&L=confocalmicroscopy&D=0&P=18718). In practice, δ is only known within certain bounds, and often even the assumption of a single- exponential intensity decay does not hold. The reasons are the following: (i), whereas λ and n3 are identified, neither n1 nor θ are known with much precision. For a biological cell, n1 varies on a microscopic length-scale [14] modifying both θc and δ, Fig. 1C. But also θ is only known with a certain accuracy (adjustment accuracy) and precision (beam divergence). These unknowns translate into an uncertainty and range of penetration depths δ(θ), respectively. (ii), even with all parameters known, the calculated penetration depth is just that: theoretical, because microscope- and sample-generated non-evanescent light modifies the intensity decay that will no longer be a simple exponential. For prismless TIRF [15] several reports have shown that excitation light distribution is best described by the superposition of a rapidly decaying and a long-range component that is fairly independent of θ [16; 17; 18]. The effect of this long-range component is paradoxical: while most of the excited fluorescence is due to EW-excitation for θ just above θc, for very high incidence angles (that normally should produce better optical sectioning) non-evanescent light dominates. (iii), background comes from stray excitation from inside the microscope objective and different optical surfaces [17]. In the critical illumination scheme used in multi-angle TIRF, any scatter on the scanning mirrors (or in any conjugate sample plane) is imaged into the sample plane, Figs. 2A and 2B, (a). Although this stray excitation can be quantitated [16; 17] such measurements are not routine; arXiv - presubmission 5 Oheim et al. (2019) TIRF calibration (iv), protein-rich adhesion sites or near-membrane organelles like lysosomes or dense-core granules have a higher RI than the surrounding cytoplasm. The spatial inhomogeneity in n1(x, y, z) does not only affect δ directly (eq.1), but it also produces non- evanescent light, Fig. 2B, panel (b). Scattering occurs predominantly in forward direction, into angles close to the original propagation direction of the EW [17; 19; 20]. For even higher local RI ≥ n1, light can be refracted, generating intense beams in EW propagation direction, as observed in chromaffin cells packed with secretory granules [19], Fig. 2B, (c). In view of these difficulties, authors and microscope manufacturers have resorted to calculating δ(θ) using eq.(1) -- see https://imagej.nih.gov/ij/plugins/tirf/index.html for a popular ImageJ plug-in -- leading to overly optimistic if not unrealistic statements of light confinement at the reflecting interface. The interpretation of TIRF intensities in terms of fluorophore concentration changes, fluorophore axial distances or single-particle trajectories is thus often flawed by large error bars or, worse, simply wrong. Azimuthal beam spinning does not improve axial confinement At least for illumination non-homogeneities there is remedy. A straightforward solution is 2- photon EW excitation [20; 21; 22]. Another is azimuthal beam scanning [23; 24; 25]: as scattering is directional, varying the EW propagation direction scrambles both propagation and scattering directions and reduces the flare in any given direction, Fig. 2C. 'Spinning' TIRF (spTIRF) or incoherent ring illumination [26] reduces interference fringes and illumination non-uniformities but it does not change the problem of the presence of non- evanescent excitation light [17] as it only redistributes and equalizes intensities, Fig. 2C. spTIRF is increasingly being used [17; 27; 28; 29; 30; 31; 32; 33; 34; 35] and commercial systems are available (e.g., from TILL: 'Polytrope' [36], Leica [37], Roperscientific iLas [34]), however, the bulk of published TIRF images has been acquired with unidirectional illumination. (figure 2 Illumination imperfections and azimuthal beam-scanning TIRF close to here) Why calibrating penetration depths at all? There are a number of biophysical techniques that do not just aim at background rejection but make quantitative use of TIRF intensities, either for localizing fluorophores or for measuring axial concentration profiles: arXiv - presubmission 6 Oheim et al. (2019) TIRF calibration • VA-TIRF. Variable-angle TIRF (VA-TIRF) is a technique that uses systematic variations of θ for reconstructing axial fluorophore distributions, Fig. 3A. For a given axial fluorophore profile C(z) the fluorescence F(θ) displays a characteristic shape. With θ known and F(θ) measured, C(z) can be obtained, pixel by pixel by an inversion procedure [19; 38; 39; 40; 41; 42; 43; 44]. VA-TIRF relies on the precise knowledge of the shape of the axial intensity decay. Many studies assumed a mono- exponential intensity decay [45; 46; 47; 48; 49], which perhaps holds for prism-type TIRF but seems overly optimistic for objective-type TIRF [16; 17; 33]. • TIRF-colocalization. Knowledge of the axial intensity decay is mandatory for multi- color excitation, Fig. 3B. As the penetration depth scales linearly with λ (eq.1), one can compensate the λ-dependence of δ by adjusting θ to maintain a constant probe volume in different color channels [34; 35; 50; 51]. The same applies to controls in TIR-FRET [31] with direct and FRET excitation of donor fluorescence. • TIRF-FRAP [52], Fig. 3C, uses the combination of localized EW-photobleaching and TIRF imaging of the fluorescence recovery after photobleaching to study near- interface fluorophore mobility. The interpretation of FRAP data relies on the known probe volume. δ(θ) calibration is even more stringent for bleaching- or photoswitching-based axial super-resolution measurements based upon consecutive VA-TIRF imaging of deeper and fluorescence deletion in more proximal layers [53]. Analogous arguments hold for TIRF photoactivation and photoswitching experiments, which include the growing group of PALM/STORM super-resolution studies, as well as optogenetic activation using EW-illumination [54]. • TIR-FCS. TIRF correlation spectroscopy [55; 56; 57] follows the same principles as confocal-spot FCS but gains sensitivity and surface selectivity from the additional excitation confinement, particularly when combined with confocal-spot TIRF excitation [58; 59]. • Even for less specialized TIRF applications a known penetration depth is a pre- requisite for reproducibility and comparing data among experiments, laboratories and publications. (figure 3 Quantitative uses of TIRF close to here) arXiv - presubmission 7 Oheim et al. (2019) TIRF calibration In the sequel, we review techniques for measuring θ and then calibrating δ(θ) and we discuss their respective benefits and problems. We also comment on supercritical angle fluorescence (SAF) microscopy [60; 61], or "evanescence in emission" [4; 62; 63] and how it can be used either as an axial nanoscale ruler [64; 65] or be combined with TIRF to achieve a better near-membrane confinement than obtained with TIRF alone [33]. Finally, we describe a simple yet effective way for calibrating the true optical sectioning of the microscope, based on the acquisition of combined TIRF and EPI z-image stacks of a thin 3- D sample of sub-diffraction beads embedded in a low-melting point agarose gel that mimics the refractive index of the cytoplasm [66]. Measuring the polar beam angle Most techniques for calibrating δ rely on determining θ and measuring the intensity resulting from a known axial fluorophore distribution F(z). A minimal θ calibration can be obtained by measuring the positions of the setscrew or stepper motor at three characteristic points: (i), normal incidence, θ = 0, epifluorescence (EPI), (ii), the critical angle θc and, (iii), the limiting angle θNA of the objective. Identified either from sample-plane [20] or back- focal plane (BFP) images [17; 25; 48; 67], Abbe's sine condition, r = fobj n1 sinθ, is used to extrapolate to intermediate values. r is the radius in the exit pupil plane (objective BFP), and f is the focal length of the objective, fobj = fTL/M, with fTL being the focal length of the manufacturer tube lens, and M is the objective transversal magnification, Fig. 4A. For a VA- or spTIRF setup, one can substitute r using the focal length of the focusing lens, fFL, and the substrate refractive index n1 to obtain 𝜃=sin!! !!"!!"#∙!"#!∙!!!! , (Eq.2) and k is a constant (°/V) characteristic for the scanner used, and Uθ is the voltage applied to the polar axis of the scanner. Alternatively, one can measure θ in the sample plane, with the advantage that coverslip tilt and beam misalignment are accounted for. In the 'lateral-displacement' technique, the laser is set to an oblique angle and the coverslip is topped with a drop of dye solution or a thin flurophore film. Defocus produces a lateral movement of the fluorescent spot that can be traced from fitting a 2-Gaussian with the elliptical cross-section of the arXiv - presubmission 8 Oheim et al. (2019) TIRF calibration beam, Fig. 4B. Fitting a straight line with its center position yields arcsin(θ) [18; 68]. Alternatively, for the beam not to suffer TIR, the glass/air surface can be made transmissive by an oil-coupled external prism [25; 35] or a solid-immersion lens (SIL) [17], Fig. 4C, and the (refracted) beam is projected to the ceiling or wall. In a variant of this triangulation technique, the back-reflected beam is picked up and projected onto a quadrant photodiode to determine θ from the center-of-mass of a Gaussian fitted with the beam profile [69]. Similarly, taking out the emission filter permits to see the back-reflected light on a BFP image, and calculating θ = arcsin[r/(n1 fobj)] [67], Fig. 4D. (figure 4 Polar beam-angle calibration close to here) Intensity-based measurements of evanescent-wave penetration depths Once a look-up table for θ has been generated, we can adjust and estimate δ(θ). This involves localizing the reflecting interface and measuring the fluorescence (or some other variable) for different fluorophore heights. This procedure is repeated for several incidence angles θ and the obtained curve compared with the calculated one (fig. 1B). a) Calibration samples with a known axial fluorophore profile. Fig. 5 shows examples of such samples, including the "raisin cake", a random, sparse 3-D distribution of sub- diffraction fluorescent microspheres in an index-matched gel, Fig. 5A. The acquisition of a z-stack of images can localize these beads with respect to the reflecting interface [66]. Alternatively, point emitters can be fixed to the surface of an oblique microscope slide, Fig. 5B, [34], a large convex [70] or concave lens [67]. Equivalently, one can use the contour of an index-matched dye-coated large-diameter fluorescent bead, or an unlabeled bead embedded in dye-containing medium [16], Fig. 5C. An elegant variant is the use of a tilted fluorescently labeled microtubule [71]. All the above test samples have in common the requirement for an evenly lit field of view (see [17] for the limits of this approximation) and they all need z-scans to locate the fluorophores with respect to the reflecting interface. This adds a complication for objective- type TIRF as the reflected beam displaces laterally upon focusing (see above), and the z- dependence of the PSF (detection volume) and z-dependent spherical aberrations require for a correction, see [71] and our calibration protocol, below. arXiv - presubmission 9 Oheim et al. (2019) TIRF calibration b) Semi-infinite dye layers. Simpler is the use of VA-TIRF and a thick, d>>δ(θ), homogenous fluorescent sample [20; 40] to estimate the effective penetration depth from the variation of the cumulative fluorescence, Fig. 5D. Assuming a two-component axial decay, i.e., the sum of mono-exponentially decaying EW with a decay length δ(θ) and a long-range component with D >> δ [16; 18] we express the measured fluorescence as 𝐹𝑧 =𝑏+𝐴∙𝑒!!/!(!)+𝐵∙𝑒!!/! (Eq.3) Here, we assumed that D to be only slowly varying with θ and the offset b negligible. After integration over z in the bounds [0, ∞], eq.3 yields a linear dependence of the measured fluorescence on δ(θ), 𝐹!"!=𝐴∙𝛿𝜃 ∙ 1−𝑒! !!! +𝐵∙𝐷∙ 1−𝑒!!! ≈𝐴∙𝛿𝜃 +𝐵, (Eq.4) because the second term is an angle-independent offset. If eq.4 is normalized for the θ- dependence of the incident intensity at the interface, I0(z=0), e.g., by recording F0(θ) of a thin fluorophore film at the interface [20; 42; 72], then the implicit θ-dependence of A and B is cancelled out. For the integral to solve as shown one has to assume that the θ-dependent term of the intensity decay follows a mono-exponential. c) Fluorescence correlation spectroscopy (FCS). Another way to determine the axial confinement is TIR-FCS [55; 56; 57]. The intensity fluctuations resulting from single molecules moving in and out of the excitation volume are being used to estimate its size, Fig. 5E. Assuming a mono-exponential intensity decay, one can explicitly solve the autocorrelation function and back-calculate the dye concentration, the bulk diffusion coefficient and the excitation volume [55]. Complications arise from dye adsorption and hindered mobility at the interface, from the distance-dependent fluorescence collection efficiency [40] as a consequence of the change in the fluorophore radiation pattern for interface-proximal dipoles (see below) and of a non-exponential axial intensity decay. d) Single-spot measurements. Other approaches measure δ(θ) only in one point, either by sampling the local EW intensity with a thinned optical fiber tip connected to a photodetector [73], or by measuring the fluorescence generated by a sub-resolution fluorescent bead fixed at the tip of an AFM cantilever [74; 75; 76] or the tip of a micropipette [45]. Brutzer et al. arXiv - presubmission 10 Oheim et al. (2019) TIRF calibration used a four-arm DNA junction as a nanomechanical translation stage to propel a single fluorescent quantum dot through EW field [77], Fig. 5F. A similar strategy uses a combination of magnetic tweezers and a supercoiling DNA together with a fluorescent nanodiamond-labeled magnetic bead and surface-immobilized fluorescent nanodiamonds as a reference for z = 0 [78]. Alternatively, fluorophores at different distances z could be obtained with static 3-D scaffolds, e.g., tetrahedral DNA-origami fluorescent rulers of different dimensions [79; 80; 81]. (figure 5 close to here) e) Index-matched polymer steps. Many of the above approaches either perturb the EW by the presence of RI boundaries (i.e., the edge of a glass slide, a lens touching the interface, or a fiber tip), they require tedious sample preparation, or they are not applicable in the aqueous environment of biological TIRF. Taniguchi's group introduced a calibration slide featuring steps of different nanometric heights fabricated from a non-fluorescent polymer, the RI of which matched that of water. This spacer staircase was topped with sub-diffraction fluorescent microspheres generating fluorescent steps at different distances from the reflecting interface [82; 83], Fig. 6A, left. The Schwille lab recently proposed a similar strategy using - instead of combining several nanofabribation techniques as in the Unno papers - a simpler dip-coating method. They topped their staircase polymer with AlexaFLuor488 solution and measured the EW-excited fluorescence as a function of step height [18], reminiscent of the "infinite dye layer" technique, Fig. 6A, right. Corroborating earlier work, both studies confirmed that the axial intensity profile contained both evanescent and non-evanescent contribution and that the effective penetration depth exceeded the calculated one. A yet different approach uses thin films of optically transparent polymers on glass substrates wrapping a nanometric dye layer. Fig. 6B illustrates such a 'sandwich' consisting of a non-fluorescent spacer layer that onto which J-aggregates were electrostatically adsorbed, resulting in a ~2-nm thin, homogenous emitter layer, Fig. 6B, inset. This layer was finally spin cast with another, 5-µm thick, polymer layer [84]. The advantage of this multi-layer test sample is that it produces a thin, controlled and homogenous dye distribution at a precisely controlled distance from the reflecting interface. With several of such test samples, plotting the collected fluorescence as a function of the fluorophore distance Δ allows the measurement of the axial effective intensity decay, Fig. 6C. arXiv - presubmission 11 Oheim et al. (2019) TIRF calibration law. The selective detection of (figure 6 close to here) Other descriptors than intensity Until now, we have focused on techniques for calibrating the EW that relied on fluorescence intensity measurements from test samples having a known axial fluororphore profile. However, other parameters of fluorescence can be used, too. Fluorophore radiation pattern. Fluorophores change their radiation pattern when they approach a RI boundary closer than ~λ because the evanescent component of their radiation can couple to the interface, become propagative [85; 86; 87] and detectable in the far field at angles >θc 'forbidden' by Snell's this 'super- critical angle fluorescence' (SAF) [62; 88; 89] features a similar surface selectivity and background suppression of TIRF whilst not requiring EW excitation. As SAF probes the same sub-λ length-scale, it can be used in conjunction with EW-excitation for achieving a better surface selectivity than TIRF alone [33]. Fourier-plane imaging conveniently measures the radiation pattern [90]. The BFP image has other benefits for calibrated TIRF microscopy: it allows determining the cell's near-membrane RI, n1, from the radius at which the transition between SAF and undercritical angle fluorescence (UAF) occurs [14; 63; 91]. The SAF/UAF intensity ratio is proportional to fluorophore height [4; 63; 64; 91] a feature, which can be used for combined axial fluorophore localization and penetration-depth calibration by plotting the total fluorescence SAF+UAF vs. (SAF/UAF) [84] (Fig. 6C). Fluorescence lifetime, τ is a measure of how long the molecule stays in the excited state after absorption of a photon. τ is one over the sum of the radiative and non-radiative decay rates. Fluorescence lifetime oscillations and shortening are observed in the presence of a metal coating [92; 93], a near-field probe tip [94] or a metal nanoparticle [95]. For the sub-λ distances relevant here, non-radiative energy transfer from the excited molecule to the metal offers an alternative decay path for excited-state relaxation. Distance-dependent surface quenching by a thin gold layer was used to measure fluorophore heights of dye-labeled arXiv - presubmission 12 Oheim et al. (2019) TIRF calibration microtubules [96] on the basis of the model by Chance, Prock, and Silbey (CPS model) that relates τ to the fluorophore height z [92]. While the distance-dependent lifetime quenching is strong for metal, only a small effect is observed on bare glass [86]. Also, other factors than surface distance interfere with molecular lifetimes, including the orientation [97], RI [98], solvent polarity, viscosity and by complex formation and collisions in the presence of nearby quenchers. Other, more exotic ways for axially localizing particles and calibrating δ(θ) include interferometric, PSF-engineering-based, axially-structured illumination-based or multi-plane based axial super-resolution techniques (reviewed in [99]), but these generally require considerably more complex instrumentation. Yet other approaches use not light but other physical parameters that are modified by the presence of an interface, e.g., the anisotropic diffusion near an interface [100] or the flow velocity gradient of particles moving under the influence of a rotating disk [101] to calibrate the EW decay. In summary, experimenters dispose of a host of observables for estimating the axial intensity decay at or near a dielectric interface. In spite of a common quest for quantitative TIRF imaging, neither a consensual test sample, nor common metrics, nor an industry standard have emerged for calibrating evanescent-wave excited fluorescence intensities. Instead, different labs have come up with their own customized solutions, making the comparison among studies difficult, if not impossible. In the now dominant prismless objective-type TIRF, the quantitative interpretation of TIRF intensities is problematic due to the co-existence of a localized, evanescent and a non-evanescent, long-range excitation component that has consistently been observed using different protocols [16; 17; 18]. Furthermore, at the same beam angle and beam diameter, the same TIRF objective, used in conjunction with different illumination optics, will perform differently. It would thus seem important that the field agrees on a protocol for testing, evaluating and quantifying axial confinement in EW techniques. A simple protocol for interpreting, comparing, and sharing TIRF data A reference standard for calibrating the effective probe depth in TIRF and SAF microscopies should meet the following requirements, it arXiv - presubmission 13 Oheim et al. (2019) TIRF calibration (i) (ii) (iii) (iv) (v) should work in an aqueous environment mimicking the cytoplasm; specifically, it should not perturb the EW by introducing objects of a different refractive index, should not require modifications to existing TIRF and SAF setups, should be easily transposable from one lab to the other, be stable or, alternatively, easily and reproducibly to fabricate, should mimic the conditions at the dielectric interface typically encountered in biological TIRF microscopy, i.e., it should reproduce the refractive index of the intracellular environment, and, should be compatible with different TIRF geometries to permit comparisons among setups and experiments. We propose the 'raisin cake' method for measuring the effective penetration depth in a sparse 3-D fluorophore distribution. The method is based on commercially available fluorescent microspheres located on the glass/water interface and at different axial positions in a transparent, index-matched agarose gel [102]. It does not require any specific equipment other than that available in any wet lab. However, in as much as the calibration procedure involves the acquisition of z-stacks of images, a precise focus control for accurate and nanometric focusing is needed. 100-nm TetraSpecksTM beads were deposited on a coverslip and at different axial distances from the coverslip by embedding them in a low-density agarose solution containing sucrose to increase the refractive index to 1.374 (see Appendix for details). We first took an epifluorescence (EPI) image stack across the sample in which the individual image planes were spaced by Δz = 10 nm, covering a range from -1.0 µm below to +2 µm above the reflecting interface. Next we acquired for the focal position corresponding to the plane in which the TIRF image was sharp (z0), one EPI and three TIRF images, varying the polar beam angle θ between each TIRF acquisition. We used spTIRF to even out illumine- tion heterogeneities, Fig. 7A. Acquisitions were realized with autofocus feedback to avoid focal drift during recordings. We determined the axial location of each bead by measuring the background-subtracted intensity in a 3×3 pixel region of interest (ROI) centered on each bead and plotted it against z. Fig. 7B shows examples of two beads located at different axial positions. We fitted Gaussian distributions with the axial fluorophore profiles and chose the peak location of the Gaussian as the bead position. Similar to the sub-diffraction xy- localization precision in PALM or dSTORM, our method allowed us to determine bead arXiv - presubmission 14 Oheim et al. (2019) TIRF calibration positions with 15-nm precision, well below the depth-of-field of the objective. 100-nm diameter beads were found to give more consistent results that 200-nm beads. However, even for the smaller beads the axial intensity profile did not always follow a Gaussian distribution. In some cases impurities or microbubbles in the agarose, low signal-to-noise or the superimposition of 2 beads distorted the profile, Fig. 7B, right. Other errors arise from beads that are directly attached to the coverslip because the vicinity to the glass/medium interface distorts the axial intensity profile due to spherical aberration at the RI-boundary, Fig. 7B, right. We therefore eliminated all beads that displayed a non-Gaussian profile from analysis. Finally, the peak of the Gaussian distribution on the EPI z-stack was taken as the intrinsic fluorescence of the beads, Fig. 7C), and used to normalize their fluorescence. (figure 7 Raisin-cake calibration close to here) The EPI image acquired at z0 allowed us to assess the influence of the depth-of-focus of the objective on the fluorescence intensity of beads remote from the interface. Keeping the focus at z0 , the EPI intensity of the off-focus beads decayed with distance to the focal plane with a length constant of 481 ± 166 nm, Fig.7D. The same intensity envelope will also modulate the bead intensities in TIRF and be convoluted with the EW excitation intensity decay. Indeed, axial intensity decay of the beads upon TIRF was best described by the superposition of a long-range and short-range exponential decay (Eq.3). We set the long- range component to the earlier measured EPI decay (D = 481 ± 166 nm) whereas the short- range component was modulated by changing θ. Focusing at the TIRF-illuminated layer (z0) we extracted the fluorescence intensity of all beads in the field of view at different polar angles, ranging from just above θc to θNA. For each bead, the obtained fluorescence intensity was normalized to the intrinsic (in-focus) EPI fluorescence intensity of the bead and plotted against its z-position and a double exponential fitted with the ensemble of data points. We took the short-range component as the effective penetration depth δ(θ). At θ = 67.5°, 70.0 and 72.5°, δ(θ) was (139 ± 20) nm, (109 ± 16) nm, and (91 ± 13) nm, respectively, close to the calculated penetration depths of 145, 111 and 94 nm, respectively, given by the software by the iLAS module (Gataca systems, France). At θ = 65°, close to θc = asin(1.375/1.52) =64.8°, fitting a single- or double exponential with the measured fluorescence intensity decay similarly resulted in δ(65°) = 367 ± 217 or 349 ± 118 nm, respectively, indicating that non-evanescent excitation light indistinguishable from EPI dominated and impaired the arXiv - presubmission 15 Oheim et al. (2019) TIRF calibration penetration depth estimation, compared to a calculated δcalc(65°) = 308 nm. Thus, similar to earlier reports [16; 25; 33 Niederauer, 2018 #67] [18], we find that supercritical-angle illumination through the periphery of a high-NA objective produces a mix of evanescent and EPI illumination. For beam angles of 72.5°, 70.0 and 67.5°, the fractional weight B/(b +A+B) from eq.(3) of non-evanescent excitation light amounted to 13, 15.3, and 15.5%. However, to which degree of evanescent vs. non-evanescent light a fluorophore is exposed to, will depend on its very distance from the reflecting interface and, paradoxically, the contribution of non-evanescent light much exceeds the above percentages for z >> 0. Our technique permits a reliable measurement of the effective axial confinement at the reflecting interface. However, all depends on the axial-localization accuracy of the beads, which in turn is a function of signal. The same protocol is applicable to fixed cells, in which small organelles are labeled (not shown). In this case, the density of labeled organelles has to be low to avoid the superimposition of fluorescent structures. Furthermore, the object size and their fluorescent intensity have to be fairly homogenous to avoid artifacts (see Appendix). Conclusions Prism-based TIRF can provide 'cleaner' and more uniform illumination as well as a higher-accuracy angular positioning, therefore it should be applied where the acquisition of an accurate intensity profile is of importance. Point-scanning TIRF (albeit slower) has slightly better lateral resolution than wide-field TIRF and can be combined with SAF detection to improve axial confinement. Independent of the very geometry used, variety of EW-calibration technique exists, among which we recommend the simple, inexpensive and reproducible 'raisin-cake' technique. • first, calibrate the polar beam angle θ, either by ray tracing through the illumination optical path or via triangulation after the passage of the beam through the objective; • Use SAF refractometry with the same fluorophores you plan to use for TIRF imaging to get an idea of the average refractive index, <n1>, surrounding the label; • Use polar beam angles not to close to the critical angle; • Use θ-sweeps to study how different angles affect the TIRF image; • Use azimuthal beam spinning TIRF to homogeneously illuminate the scene; arXiv - presubmission 16 Oheim et al. (2019) TIRF calibration • Calibrate the effective penetration depth using the 'raisin cake' technique; • Systematically report θ ± Δθ, <n1 ± Δn1>, and the estimated δ ± Δδ; • Combine TIRF excitation and vSAF detection for better optical sectioning; • Discuss, how the uncertainty in δ and TIRF intensities affects your conclusions. APPENDIX A: TIRF test-sample protocol Sample preparation. Small droplets of a 1:4 dilution in water of 0.1-µm diameter TetraspeckTM beads (Invitrogen) were deposited on a glass coverslip (BK-7, e.g., FisherScientific or Marienfeld) and allowed to dry, Fig. S1A. The bead surplus was flushed with water. In parallel, a solution of 0.5% low-melting point agarose solution (seaplaque GTG agarose, Lonza) containing 842 mM sucrose was prepared. Sucrose increased the RI of the solution to 1.375 to mimic the refractive index of cells [103]. The solution was heated to ~80°C in a water bath to melt the agarose. It was then kept at constant ~50°C. 5 µl of sonicated beads were transferred to a pre-heated Eppendorf tube to which 15 µl agarose/sucrose solution was added and thoroughly triturated with a preheated pipette. The now 20 µl of solution was pipetted onto the coverslip at the same location, topping the beads that were previously attached to the glass (Fig. S1B). The agarose was allowed to polymerize. To speed up this step the coverslip can be placed on an ice-cold surface. Then, 1-2 ml agarose/sucrose solution (not containing beads) was slowly deposited on top of the bead-containing drop to achieve a thick layer of RI = 1.375. To obtain good results, it is important that this solution is maintained at a temperature just above the gelling point of agarose (~30°C) so that the small bead-containing drop already on the coverslip does not re- melt when the large volume of agarose/sucrose solution is added. Setup. The penetration depth was assessed on a commercial multi-angle objective-type TIRF setup (Visitron Systems, Puchheim, Germany) built around an IX83 inverted microscope equipped with an autofocus module, a UAPON100XOTIRF NA-1.49 objective (all Olympus). Precise focusing was achieved with the motorized focus (Δz = 10 nm). The beam of a 488-nm 100-mW laser was directed into the iLAS2 beam scanning system (Gataca Systems, France) allowing (θ,φ) scans. Fluorescence was collected through the same objective, a ZET405/488/561/640rpc multi-band emission filter (both from Semrock) and detected on an a ZT405/488/561/640rpc multi-band extracted with dichroic, arXiv - presubmission 17 Oheim et al. (2019) TIRF calibration Evolve-EM515 EMCCD camera (Photometrics). All setup components were controlled by VisiView (Visitron Systems GmbH). The resulting pixel size in the sample plane was 160 nm. Typical integration times were 100 ms at gain 1 and an EM gain of 200 abitrary units. Data analysis Images were analyzed with imageJ (Rasband, W.S., ImageJ, U. S. National Institutes of Health, Bethesda, Maryland, USA, http://imagej.nih.gov/ij/, 1997-2014). Fitting and further analysis of the data was performed with IGOR PRO software (Wavemetrics, Lake Oswego, OR, USA). Acknowledgements Recent work in our labs related to TIRF and SAF microscopy was financed by the Agence Nationale de la Recherche (ANR-10-INSB-04-01, grands investissements FranceBioImaging, FBI, to MO), and a Chaire d'Excellence Junior Université Sorbonne Paris Cité (USPC) to MB, and the Université Paris Descartes (invited professorship during the academic year 2017-18, to AS). MO and AS acknowledge support from a French-Israeli CNRS-WIS ImagiNano LIA grant. UB acknowledges funding from the Deutsche Forschungsgemeinschaft (DFG, CRC894). The Oheim lab is a member of the C'nano IdF and Ecole de Neurosciences de Paris (ENP) excellence clusters for nanobiotechnology and neurosciences, respectively. Author contributions MO, UB, MB, AS, and AW performed experiments, MB, UB and AW prepared samples, MO, UB and AS analyzed the data, MO wrote the manuscript with contributions from all authors. All authors have given their approval to the final version of the manuscript. Conflicting interest statement The authors declare no conflict of interest. arXiv - presubmission 18 TIRF calibration Oheim et al. (2019) List of abbreviations DNA EPI EGFP EW FCS FRAP NA PALM ROI SAF SIL SIM SMD SOFI SPR spTIRF STED STORM TIR(F(M)) VA-TIRF - - - - - - - - - - - - - - - - - - - - desoxy-ribonucleic acid epifluorescence enhanced green fluorescent protein evanescent wave fluorescence correlation spectroscopy fluorescence recovery after photobleaching numerical aperture photoactivation localization microscopy region of interest supercritical angle fluoresence solid immersion lens structured illumination microscopy single molecule detection superresolution optical fluctuation imaging surface plasmon resonance spinning TIRF stimulated emission depletion stochastic optical reconstruction microscopy total internal reflection (fluorescence (microscopy)) variable-angle TIRF arXiv - presubmission 19 Oheim et al. (2019) TIRF calibration REFERENCES [1] H. Schneckenburger, Total internal reflection fluorescence microscopy: technical innovations and novel applications. Curr Op Biotechnol 16 (2005) 13-18. [2] K.N. Fish, Total internal reflection fluorescence (TIRF) microscopy. Curr. Prot. Cytometry (2009) 12.18. 1-12.18. 13. [3] A.L. Mattheyses, S.M. Simon, and J.Z. Rappoport, Imaging with total internal reflection fluorescence microscopy for the cell biologist. J. Cell Sci. 123 (2010) 3621-3628. [4] D. Axelrod, Evanescent excitation and emission in fluorescence microscopy. Biophys. J. 104 (2013) 1401-9. [5] M.L. Martin-Fernandez, C.J. Tynan, and S.E. Webb, A 'pocket guide' to total internal reflection fluorescence. J. Microsc. 252 (2013) 16-22. [6] N.S. Poulter, W.T. Pitkeathly, P.J. Smith, and J.Z. Rappoport, The physical basis of total internal reflection fluorescence (TIRF) microscopy and its cellular applications. Methods Mol Biol 1251 (2015) 1-23. [7] M. Oheim, TIRF (Total Internal Reflection Fluorescence). eLS (2016) DOI: 10.1002/9780470015902.a0022505. [8] L.J. Young, F. Ströhl, and C.F. Kaminski, A guide to structured illumination TIRF microscopy at high speed with multiple colors. JoVE (2016). [9] K.-F. Giebel, C. Bechinger, S. Herminghaus, M. Riedel, P. Leiderer, U. Weiland, and M. Bastmeyer, Imaging of cell/substrate contacts of living cells with surface plasmon resonance microscopy. Biophys. J. 76 (1999) 509-516. [10] V. Yashunsky, V. Lirtsman, M. Golosovsky, D. Davidov, and B. Aroeti, Real-time monitoring of epithelial cell-cell and cell-substrate interactions by infrared surface plasmon spectroscopy. Biophys. J. 99 (2010) 4028-4036. [11] T. Son, J. Seo, I.-H. Choi, and D. Kim, Label-free quantification of cell-to-substrate separation by surface plasmon resonance microscopy. Optics Commun. 422 (2018) 64-68. [12] W. Reichert, and G. Truskey, Total internal reflection fluorescence (TIRF) microscopy. I. Modelling cell contact region fluorescence. J Cell Sci 96 (1990) 219-230. [13] D. Axelrod, Fluorescence excitation and imaging of single molecules near dielectric‐coated and bare surfaces: a theoretical study. Journal of microscopy 247 (2012) 147-160. [14] M. Brunstein, L. Roy, and M. Oheim, Near-membrane refractometry using supercritical angle fluorescence. Biophys. J. 112 (2017) 1940-1948. [15] A.L. Stout, and D. Axelrod, Evanescent field excitation of fluorescence by epi-illumination microscopy. Appl. Opt. 28 (1989) 5237-5242. [16] A.L. Mattheyses, and D. Axelrod, Direct measurement of the evanescent field profile produced by objective-based total internal reflection fluorescence. J. Biomed. Opt. 11 (2006) 014006- 014006-7. [17] M. Brunstein, M. Teremetz, K. Hérault, C. Tourain, and M. Oheim, Eliminating unwanted far- field excitation in objective-type TIRF. Part I. Identifying sources of nonevanescent excitation light. Biophys. J. 106 (2014) 1020-1032. [18] C. Niederauer, P. Blumhardt, J. Mücksch, M. Heymann, A. Lambacher, and P. Schwille, Direct characterization of the evanescent field in objective-type total internal reflection fluorescence microscopy. Opt. Express 26 (2018) 20492-20506. [19] A. Rohrbach, Observing secretory granules with a multiangle evanescent wave microscope. Biophys. J. 78 (2000) 2641-2654. [20] F. Schapper, J.T. Gonçalves, and M. Oheim, Fluorescence imaging with two-photon evanescent wave excitation. Eur. Biophys. J. 32 (2003) 635-643. arXiv - presubmission 20 Oheim et al. (2019) TIRF calibration [21] M. Oheim, and F. Schapper, Non-linear evanescent-field imaging. J. Phys. D: Appl. Phys. 38 (2005) R185. [22] R.S. Lane, A.N. Macpherson, and S.W. Magennis, Signal enhancement in multiphoton TIRF microscopy by shaping of broadband femtosecond pulses. Opt. Express 20 (2012) 25948- 25959. [23] A.L. Mattheyses, K. Shaw, and D. Axelrod, Effective elimination of laser interference fringing in fluorescence microscopy by spinning azimuthal incidence angle. Microsc. Res. Tech. 69 (2006) 642-647. [24] R. Fiolka, Y. Belyaev, H. Ewers, and A. Stemmer, Even illumination in total internal reflection fluorescence microscopy using laser light. Microsc. Res. Tech. 71 (2008) 45-50. [25] M. van't Hoff, V. de Sars, and M. Oheim, A programmable light engine for quantitative single molecule TIRF and HILO imaging. Opt. Express 16 (2008) 18495-18504. [26] D. Axelrod, Cell-substrate contacts illuminated by total internal reflection fluorescence. J Cell Biol 89 (1981) 141-145. [27] D.S. Johnson, J.K. Jaiswal, and S. Simon, Total internal reflection fluorescence (TIRF) microscopy illuminator for improved imaging of cell surface events. Curr. Prot. Cytometry (2012) 12.29. 1-12.29. 19. [28] S. Abdelhady, S.S. Kitambi, V. Lundin, R. Aufschnaiter, P. Sekyrova, I. Sinha, K.T. Lundgren, G. Castelo-Branco, S. Linnarsson, R. Wedlich-Soldner, A. Teixeira, and M. Andang, Erg channel is critical in controlling cell volume during cell cycle in embryonic stem cells. PLoS One 8 (2013) e72409. [29] A.H. Crevenna, N. Naredi-Rainer, A. Schonichen, J. Dzubiella, D.L. Barber, D.C. Lamb, and R. Wedlich-Soldner, Electrostatics control actin filament nucleation and elongation kinetics. J. Biol. Chem. 288 (2013) 12102-13. [30] P.-I. Ku, Anna K. Miller, J. Ballew, V. Sandrin, Frederick R. Adler, and S. Saffarian, Identification of Pauses during Formation of HIV-1 Virus Like Particles. Biophys. J. 105 (2013) 2262-2272. [31] J. Lin, and A.D. Hoppe, Uniform total internal reflection fluorescence illumination enables live cell fluorescence resonance energy transfer microscopy. Microsc. Microanal. 19 (2013) 350- 359. [32] P.-I. Ku, M. Bendjennat, J. Ballew, M.B. Landesman, and S. Saffarian, ALIX Is Recruited Temporarily into HIV-1 Budding Sites at the End of Gag Assembly. PLoS ONE 9 (2014) e96950. [33] M. Brunstein, K. Hérault, and M. Oheim, Eliminating unwanted far-field excitation in objective-type TIRF. Part II. Combined evanescent-wave excitation and supercritical-angle fluorescence detection improves optical sectioning. Biophys. J. 106 (2014) 1044-1056. [34] J. Boulanger, C. Gueudry, D. Münch, B. Cinquin, P. Paul-Gilloteaux, S. Bardin, C. Guérin, F. Senger, L. Blanchoin, and J. Salamero, Fast high-resolution 3D total internal reflection fluorescence microscopy by incidence angle scanning and azimuthal averaging. Proc. Acad. Sci. USA 111 (2014) 17164-17169. [35] W. Zong, X. Huang, C. Zhang, T. Yuan, L.-l. Zhu, M. Fan, and L. Chen, Shadowless- illuminated variable-angle TIRF (siva-TIRF) microscopy for the observation of spatial- temporal dynamics in live cells. Biomed. Opt. Express 5 (2014) 1530-1540. [36] J. Riedl, A.H. Crevenna, K. Kessenbrock, J.H. Yu, D. Neukirchen, M. Bista, F. Bradke, D. Jenne, T.A. Holak, Z. Werb, M. Sixt, and R. Wedlich-Söldner, Lifeact: a versatile marker to visualize F-actin. Nat. Methods 5 (2008) 605-7. [37] T. Veitinger, Controlling the TIRF Penetration Depth is Mandatory for Reproducible Results. http://www.leica-microsystems.com/science-lab/controlling-the-tirf-penetration-depth-is- mandatory-for-reproducible-results/ Online (2012). arXiv - presubmission 21 Oheim et al. (2019) TIRF calibration [38] W. Reichert, P. Suci, J. Ives, and J. Andrade, Evanescent detection of adsorbed protein concentration-distance profiles: fit of simple models to variable-angle total internal reflection fluorescence data. Appl. Spectrosc. 41 (1987) 503-508. [39] J.S. Burmeister, G.A. Truskey, and W.M. Reichert, Quantitative analysis of variable‐angle total internal reflection fluorescence microscopy (VA‐TIRFM) of cell/substrate contacts. J Microsc 173 (1994) 39-51. [40] B.P. Ölveczky, N. Periasamy, and A. Verkman, Mapping fluorophore distributions in three dimensions by quantitative multiple angle-total internal reflection fluorescence microscopy. Biophys J 73 (1997) 2836-2847. [41] M. Oheim, D. Loerke, B. Preitz, and W. Stuhmer, Simple optical configuration for depth- resolved imaging using variable-angle evanescent-wave microscopy, Optical biopsies and microscopic techniques III, International Society for Optics and Photonics, 1999, pp. 131- 141. [42] D. Loerke, B. Preitz, W. Stuhmer, and M. Oheim, Super-resolution measurements with evanescent-wave fluorescence-excitation using variable beam incidence. Journal of biomedical optics 5 (2000) 23-31. [43] K. Stock, R. Sailer, W.S. Strauss, M. Lyttek, R. Steiner, and H. Schneckenburger, Variable‐ angle total internal reflection fluorescence microscopy (VA‐TIRFM): realization and application of a compact illumination device. J Microsc 211 (2003) 19-29. [44] J. Li, W. Han, Y. Li, Y. Chen, Y. Shang, Y. Chen, and Z. Gui, Inverse problem based on the fast alternating direction method of multipliers algorithm in multiangle total internal reflection fluorescence microscopy. Appl. Opt. 57 (2018) 9828-9834. [45] M. Oheim, D. Loerke, W. Stühmer, and R.H. Chow, The last few milliseconds in the life of a secretory granule. Eur. Biophys. J. 27 (1998) 83-98. [46] C.D. Byrne, A.J. de Mello, and W.L. Barnes, Variable-angle time-resolved evanescent wave- induced fluorescence spectroscopy (VATR-EWIFS): a technique for concentration profiling fluorophores at dielectric interfaces. The Journal of Physical Chemistry B 102 (1998) 10326- 10333. [47] M. van't Hoff, M. Reuter, D.T. Dryden, and M. Oheim, Screening by imaging: scaling up single-DNA-molecule analysis with a novel parabolic VA-TIRF reflector and noise- reduction techniques. PhysChemChemPhys 11 (2009) 7713-7720. [48] M.C. Dos Santos, R. Déturche, C. Vézy, and R. Jaffiol, Axial nanoscale localization by normalized total internal reflection fluorescence microscopy. Opt. Lett. 39 (2014) 869-872. [49] M.C. Dos Santos, R. Déturche, C. Vézy, and R. Jaffiol, Topography of cells revealed by variable-angle total internal reflection fluorescence microscopy. Biophys J 111 (2016) 1316- 1327. [50] M. Oheim, and W. Stuhmer, Multiparameter evanescent-wave imaging in biological fluorescence microscopy. IEEE J Quant Elec 38 (2002) 142-148. [51] U. Becherer, T. Moser, W. Stühmer, and M. Oheim, Calcium regulates exocytosis at the level of single vesicles. Nat Neurosci 6 (2003) 846. [52] E. Pryazhnikov, D. Fayuk, M. Niittykoski, R. Giniatullin, and L. Khiroug, Unusually strong temperature dependence of P2X3 receptor traffic to the plasma membrane. Fronti. Cell. Neurosci. 5 (2011) 27. [53] Y. Fu, P.W. Winter, R. Rojas, V. Wang, M. McAuliffe, and G.H. Patterson, Axial superresolution via multiangle TIRF microscopy with sequential imaging and photobleaching. Proc. Acad. Sci. USA 113 (2016) 4368-4373. [54] D. Li, K. Hérault, E.Y. Isacoff, M. Oheim, and N. Ropert, Optogenetic activation of LiGluR‐ expressing astrocytes evokes anion channel‐mediated glutamate release. J. Physiol. 590 (2012) 855-873. arXiv - presubmission 22 Oheim et al. (2019) TIRF calibration [55] S. Harlepp, J. Robert, N. Darnton, and D. Chatenay, Subnanometric measurements of evanescent wave penetration depth using total internal reflection microscopy combined with fluorescent correlation spectroscopy. Applied physics letters 85 (2004) 3917-3919. [56] K. Hassler, M. Leutenegger, P. Rigler, R. Rao, R. Rigler, M. Gösch, and T. Lasser, Total internal reflection fluorescence correlation spectroscopy (TIR-FCS) with low background and high count-rate per molecule. Optics Express 13 (2005) 7415-7423. [57] N.L. Thompson, X. Wang, and P. Navaratnarajah, Total internal reflection with fluorescence correlation spectroscopy: Applications to substrate-supported planar membranes. J. Struct. Biol. 168 (2009) 95-106. [58] T. Ruckstuhl, and S. Seeger, Attoliter detection volumes by confocal total-internal-reflection fluorescence microscopy. Opt. Lett. 29 (2004) 569-571. [59] M. Leutenegger, C. Ringemann, T. Lasser, S.W. Hell, and C. Eggeling, Fluorescence correlation spectroscopy with a total internal reflection fluorescence STED microscope (TIRF-STED-FCS). Opt. Express 20 (2012) 5243-5263. [60] T. Ruckstuhl, and D. Verdes, Supercritical angle fluorescence (SAF) microscopy. Opt. Express 12 (2004) 4246-4254. [61] T. Barroca, K. Balaa, J. Delahaye, S. Lévêque-Fort, and E. Fort, Full-field supercritical angle fluorescence microscopy for live cell imaging. Opt. Lett. 36 (2011) 3051-3053. [62] D. Axelrod, Selective imaging of surface fluorescence with very high aperture microscope objectives. J. Biomed. Opt. 6 (2001) 6-13. [63] M. Brunstein, A. Salomon, and M. Oheim, Decoding the information contained in the fluorophore radiation pattern ACS Nano 12 (2018) 11725 -- 11730. [64] C.M. Winterflood, T. Ruckstuhl, D. Verdes, and S. Seeger, Nanometer axial resolution by three-dimensional supercritical angle fluorescence microscopy. Phys. Rev. Lett. 105 (2010) 108103. [65] J. Deschamps, M. Mund, and J. Ries, 3D superresolution microscopy by supercritical angle detection. Opt. Express 22 (2014) 29081-29091. [66] A. Quintana, C. Kummerow, C. Junker, U. Becherer, and M. Hoth, Morphological changes of T cells following formation of the immunological synapse modulate intracellular calcium signals. Cell Calcium 45 (2009) 109-122. [67] E. Soubies, S. Schaub, A. Radwanska, E. Van Obberghen-Schilling, L. Blanc-Féraud, and G. Aubert, A framework for multi-angle TIRF microscope calibration, Biomedical Imaging (ISBI), 2016 IEEE 13th International Symposium on, IEEE, 2016, pp. 668-671. [68] T.P. Burghardt, Measuring incidence angle for through-the-objective total internal reflection fluorescence microscopy. J. Biomed. Opt. 17 (2012) 126007-126007. [69] D.S. Johnson, R. Toledo-Crow, A.L. Mattheyses, and S.M. Simon, Polarization-controlled TIRFM with focal drift and spatial field intensity correction. Biophys. J. 106 (2014) 1008- 1019. [70] J.A. Steyer, and W. Almers, Tracking single secretory granules in live chromaffin cells by evanescent-field fluorescence microscopy. Biophys J 76 (1999) 2262-2271. [71] C. Gell, M. Berndt, J. Enderlein, and S. Diez, TIRF microscopy evanescent field calibration using tilted fluorescent microtubules. J. Microsc. 234 (2009) 38-46. [72] M. Guo, P. Chandris, J.P. Giannini, A.J. Trexler, R. Fischer, J. Chen, H.D. Vishwasrao, I. Rey- Suarez, Y. Wu, and X. Wu, Single-shot super-resolution total internal reflection fluorescence microscopy. Nature methods 15 (2018) 425. [73] A.J. Meixner, M.A. Bopp, and G. Tarrach, Direct measurement of standing evanescent waves with a photon-scanning tunneling microscope. Appl. Opt. 33 (1994) 7995-8000. [74] A. Sarkar, R.B. Robertson, and J.M. Fernandez, Simultaneous atomic force microscope and fluorescence measurements of protein unfolding using a calibrated evanescent wave. Proc. Acad. Sci. USA 101 (2004) 12882-12886. arXiv - presubmission 23 Oheim et al. (2019) TIRF calibration [75] J. Oreopoulos, and C.M. Yip, Combined scanning probe and total internal reflection fluorescence microscopy. Methods 46 (2008) 2-10. [76] M. Franken, C. Poelma, and J. Westerweel, Nanoscale contact line visualization based on total internal reflection fluorescence microscopy. Optics express 21 (2013) 26093-26102. [77] H. Brutzer, F.W. Schwarz, and R. Seidel, Scanning evanescent fields using a pointlike light source and a nanomechanical DNA gear. Nano letters 12 (2011) 473-478. [78] Y. Seol, and K.C. Neuman, Combined Magnetic Tweezers and Micro-mirror Total Internal Reflection Fluorescence Microscope for Single-Molecule Manipulation and Visualization, Single Molecule Analysis, Springer, 2018, pp. 297-316. [79] C. Steinhauer, R. Jungmann, T.L. Sobey, F.C. Simmel, and P. Tinnefeld, DNA origami as a nanoscopic ruler for super‐resolution microscopy. Angew. Chem. 48 (2009) 8870-8873. [80] J.J. Schmied, A. Gietl, P. Holzmeister, C. Forthmann, C. Steinhauer, T. Dammeyer, and P. Tinnefeld, Fluorescence and super-resolution standards based on DNA origami. Nat. Meth. 9 (2012) 1133. [81] R. Schreiber, J. Do, E.-M. Roller, T. Zhang, V.J. Schüller, P.C. Nickels, J. Feldmann, and T. Liedl, Hierarchical assembly of metal nanoparticles, quantum dots and organic dyes using DNA origami scaffolds. Nat Nanotech 9 (2014) 74. [82] N. Unno, A. Maeda, S.-i. Satake, T. Tsuji, and J. Taniguchi, Fabrication of nanostep for total internal reflection fluorescence microscopy to calibrate in water. Microelectronic Engineering 133 (2015) 98-103. [83] N. Unno, H. Kigami, T. Fujinami, S. Nakata, S.-i. Satake, and J. Taniguchi, Fabrication of calibration plate for total internal reflection fluorescence microscopy using roll-type liquid transfer imprint lithography. Microelectronic Engineering 180 (2017) 86-92. [84] M. Oheim, and A. Salomon, Calibration standard for evanescence microscopy. in: e.a. Centre National de la Recherche Scientifique - CNRS, (Ed.), 2018. [85] C. Carniglia, L. Mandel, and K. Drexhage, Absorption and emission of evanescent photons. J. Opt. Soc. Am. 62 (1972) 479-486. [86] W. Lukosz, and R. Kunz, Light emission by magnetic and electric dipoles close to a plane interface. I. Total radiated power. JOSA 67 (1977) 1607-1615. [87] J. Mertz, Radiative absorption, fluorescence, and scattering of a classical dipole near a lossless interface: a unified description. JOSA B 17 (2000) 1906-1913. [88] J. Enderlein, T. Ruckstuhl, and S. Seeger, Highly efficient optical detection of surface- generated fluorescence. Appl. Opt. 38 (1999) 724-732. [89] T. Ruckstuhl, J. Enderlein, S. Jung, and S. Seeger, Forbidden light detection from single molecules. Anal. Chem. 72 (2000) 2117-2123. [90] A.L. Mattheyses, and D. Axelrod, Fluorescence emission patterns near glass and metal-coated surfaces investigated with back focal plane imaging. J. Biomed. Opt. 10 (2005) 054007. [91] N. Bourg, C. Mayet, G. Dupuis, T. Barroca, P. Bon, S. Lécart, E. Fort, and S. Lévêque-Fort, Direct optical nanoscopy with axially localized detection. Nat. Photonics 9 (2015) 587. [92] R. Chance, A. Prock, and R. Silbey, Lifetime of an emitting molecule near a partially reflecting [93] K. Tews, On the variation of luminescence lifetimes. The approximations of the approximative surface. J. Chem. Phys. 60 (1974) 2744-2748. methods. J. Luminesc. 9 (1974) 223-239. [94] W.P. Ambrose, P.M. Goodwin, R.A. Keller, and J.C. Martin, Alterations of single molecule fluorescence lifetimes in near-field optical microscopy. Science 265 (1994) 364-367. [95] J. Seelig, K. Leslie, A. Renn, S. Kühn, V. Jacobsen, M. van de Corput, C. Wyman, and V. Sandoghdar, Nanoparticle-induced fluorescence lifetime modification as nanoscopic ruler: demonstration at the single molecule level. Nano letters 7 (2007) 685-689. [96] M. Berndt, M. Lorenz, J. Enderlein, and S. Diez, Axial Nanometer Distances Measured by Fluorescence Lifetime Imaging Microscopy. Nano Lett. 10 (2010) 1497-1500. arXiv - presubmission 24 Oheim et al. (2019) TIRF calibration [97] M. Kreiter, M. Prummer, B. Hecht, and U. Wild, Orientation dependence of fluorescence lifetimes near an interface. J. Chem. Phys. 117 (2002) 9430-9433. [98] S. Strickler, and R.A. Berg, Relationship between absorption intensity and fluorescence lifetime of molecules. J. Chem. Phys. 37 (1962) 814-822. [99] W. Liu, K.C. Toussaint Jr, C. Okoro, D. Zhu, Y. Chen, C. Kuang, and X. Liu, Breaking the Axial Diffraction Limit: A Guide to Axial Super‐Resolution Fluorescence Microscopy. Laser Photon. Rev. 12 (2018) 1700333. [100] S. Saffarian, and T. Kirchhausen, Differential evanescence nanometry: live-cell fluorescence measurements with 10-nm axial resolution on the plasma membrane. Biophys. J. 94 (2008) 2333-2342. [101] C. Zettner, and M. Yoda, Particle velocity field measurements in a near-wall flow using evanescent wave illumination. Exp. Fluids 34 (2003) 115-121. [102] A. Quintana, C. Schwindling, A.S. Wenning, U. Becherer, J. Rettig, E.C. Schwarz, and M. Hoth, T cell activation requires mitochondrial translocation to the immunological synapse. Proc. Acad. Sci. USA 104 (2007) 14418-23. [103] X. Liang, A. Liu, C. Lim, T. Ayi, and P. Yap, Determining refractive index of single living cell using an integrated microchip. Sensors Actuators A: Physical 133 (2007) 349-354. arXiv - presubmission 25 Oheim et al. (2019) TIRF calibration FIGURE LEGENDS FIGURE 1. Fundamentals of TIRF excitation. (A), left, prismless (objective-type) TIRF. A laser beam is focused to an eccentric position in the back-focal plane (BFP, dashed) of a high-numerical aperture (NA) objective generating a collimated beam impinging at an oblique angle θ at the dielectric interface (solid grey, n3>n1). Middle, for θ exceeding the critical angle, θc =asin(n1/n3), the beam is totally reflected at the interface and an inhomogeneous surface wave is generated in the rare medium (n1). This 'evanescent' wave (EW) propagates along the surface (the Pointing vector, S, is oriented in +x direction for a beam impinging from the left, red arrow), and its intensity decays exponentially with axial (+z) distance from the reflecting interface, right, with a length constant ('penetration depth') δ of the order of 100 nm. (B), dependence of δ on θ, for λ = 488 nm, n1 = 1.35, and for different substrate indices n3. The higher n3 the smaller the critical angle θc and the better the optical sectioning. For a typical borosilicate glass/cell-interface and a NA-1.45 objective, the maximally attainable angle θΝΑ limits the penetration depth to 73 nm (red dash). In this angle range, δ depends steeply on θ, demanding high precision and accuracy when adjusting θ. (C), dependence of θc and δ∞ on substrate index, n3. The asymptotes of the critical angle θc and limiting penetration depth δ∞ for grazing incidence (θ à 90°), respectively, decrease monotonously with n3. Red line indicates n3 = 1.52 (BK-7), as before. FIGURE 2. Azimuthal beam scanning evens out illumination imperfections. (A), layout for polar- and azimuthal beam scanning TIRF ('spinning' TIRF, spTIRF). A 2-axis scan (θ', φ') is relayed via a 4f beam compressor, BC, into an equivalent sample plane, ESP, and - via the telescope formed by the focusing lens, FL, and objective (obj) -- imaged into the sample plane (SP). Inset shows light distribution in the back-focal plane, BFP. Abbreviations: dic -- dichroic mirror, EMCCD -- electron- multiplying charged-coupled device camera, EBFP -- equivalent back-focal plane, SD -- scanning device. Due to this critical illumination any illumination imperfections in ESP' and ESP show up in SP. (B), sources of non-evanescent excitation, (a), dust on the scanning mirrors and in intermediate sample planes is directly imaged into the sample plane, resulting in illumination impurities and glare; (b) EW scattering at refractive-index (RI) boundaries produces light propagating in forward direction, modifying the effective δ(θ) across the field-of-view; (c), for shallow angles θ ≥θc, protein-rich cell adhesion sites can have a RI high enough to disrupt total reflection and generate intense beams of refracted light. (C), negative-staining experiment, in which a non-labeled BON cell was embedded in fluorescein-dextran containing extracellular saline and imaged in unidirectional TIRF (top) and spTIRF (bottom), respectively (from Brunstein et al. BJ 2014a). The bottom of the cell adhering to the coverslip excludes the extracellular dye and appears as a dark 'footprint'. Note the flare in EW propagation direction (top), which is abolished upon spTIRF (bottom). FIGURE 3. Quantitative uses of TIRF. (A), variable-angle (VA-) TIRF. Smaller θ translate into larger illumination depths δ(θ) (block arrow) allowing a topographic reconstruction of axial fluorophore profiles from a multi-θ stack. (B), multi-λ excitation. δ scales linearly with λ (block arrows). Toggling between different excitation wavelengths alters the excited volume (grey arXiv - presubmission 26 Oheim et al. (2019) TIRF calibration arrowheads), top. Adjusting θ between different-color acquisitions maintains a constant excitation volume, permitting quantitative co-localization or FRET studies at or near the basal plasma membrane, bottom. (C), TIR-FRAP, in this variant of fluorescence recovery after photobleaching, a sample is sequentially imaged, bleached and re-imaged with EW excitation. An intense pulse of evanescent light (flash) selectively bleaches the surface-proximal fluorophores. Unbleached molecules from deeper sample regions repopulate the bleached volume. TIRF-FRAP allows studying the average mobility and mobile fraction of near-membrane fluorescent species, inset. Here, the penetration depth is constant, and the illumination intensity is modulated between imaging and bleaching episodes. FIGURE 4. Incidence-angle calibration. (A), the radial position r of a focused laser spot in the BFP unambiguously determines the beam angle θ. Three characteristic points can be easily identified: (1), epifluorescence (EPI) at normal incidence, θ = 0 ó r = 0; (2) the disappearance of the refracted beam at rc, ó θc (TIR) and, (3) the disappearance of the reflected beam at rNA, when the incident beam is beyond the limiting numerical aperture of the objective θNA = arcsin(NA/n1). Intermediate values of θ(r) are interpolated from Abbe's sine condition, r = fobj n1 sin(θ). (B), lateral- displacement assay. The beam is directed at an oblique angle against an oil-coupled coverslip, coated with a thin fluorophore layer, green. Depending on the amount of defocus, the beam intersects the coverslip at different positions, and the lateral offset of the fluorescent spot on the camera image, together with knowledge of the piezo-controlled defocus permits to triangulate θ. A similar strategy is used in (C) for θ > θc. Here, an oil-coupled, index-matched solid immersion lens (SIL) prohibits TIR and couples out a refracted beam that is then projected against the wall or some ruler at large distance to triangulate θ. (D), the positional information contained in the reflected beam is used to measure r(θ) by coupling out a small percentage of the reflected intensity with a miniature semitransparent micromirror (µ) in the periphery of the objective and direct it onto a position-sensitive detector (PSD), e.g., a quadrant photodiode. FIGURE 5. Intensity-based techniques for calibrating axial intensity decays. (A), 'raisin cake', sub- diffraction fluorescent microspheres embedded in an index-matched (RI≈n1) agarose gel. (B), oblique-fluorescent-layer sample, consisting of a fluorophore-coated coverslip and spacer (of height d, typically another cover glass). In a variant, (C), the surface of a long-f lens of known curvature radius is coated with fluorescent beads and positioned on the interface and again the known axial fluorophore is used to probe the EW decay. In the 'infinitely' thick (d >> λ) dye layer approach (D), a homogenous fluorophore solution is used to measure the cumulative fluorescence at a given penetration depth, δ(θ). Upon multi-θ sweeps, depending on how far the EW reaches in the fluorescent solution, the intensity changes in a predictable manner. (E), in a dilute dye solution, the intensity fluctuations resulting from the diffusion of single molecules through the EW-excited volume allow measuring the penetration depth through TIR-FCS. (F), point emitters attached to the tip of an atomic force microscope (not shown) or attached to DNA probe the EW in a single point. FIGURE 6. Refractive-index matched polymer height steps. (A), non-fluorescent polymer staircases, onto which sub-diffraction beads are drop cast to produce emitter layers at discrete distances, left, or covered with a dilute dye solution allow a tomographic reconstruction similar to arXiv - presubmission 27 Oheim et al. (2019) TIRF calibration that in Fig. 5D, right. (B), multi-layer sandwich of cell index-matched polymer spacer and capping layers (grey), separated by a thin fluorophore layer (green). (C), upon EW illumination at an angle θ, samples with different fluorophore heights Δ (top row) result in different fluorescence intensities (middle), allowing the measurement of the axial intensity decay (bottom) by integrating the total fluorescence (dashed circle). Note the excitation spots at 3- and 9-o'clock positions on the BFP images. The refractive index n1 of the polymer layers is obtained by SAF refractomertry {Brunstein, 2017 #59} from the radius at which the transition from under- to supercritical angle fluorescence occurs. The SAF/UAF ratio is directly proportional to D, {Oheim, 2018 #92}. FIGURE 7. Determining the effective penetration depth with a 'raisin-cake' test sample. (A), experimental workflow. Left, acquisition of a z-stack of images in epifluorescence (EPI, θ = 0°). Middle, EPI image taken at z0 (z := 0, red), localizing the bottom layer of beads on the cover slip. Right, corresponding TIRF image at z0 and different beam angles, θ. (B), each bead was localized from its axial intensity profile by fitting a Gaussian (black line) with the average fluorescence, F(z), measured in a 3×3 px ROI (+). Left, example bead at z = 38 nm, with F(z) well described by a Gaussian. Top images correspond to planes identified by arrows. Right, example of a distorted profile discarded from analysis, z = 44 nm. (C), Intrinsic fluorescence, i.e., bead intensity measured from the peak of the Gaussian as in (B) for each bead, vs. its in-focus position z. (D), bead fluorescence F(EPI)(z; z0) when focusing at the lowest bead layer at z0, upon EPI illumination, as a function of previously measured bead position z and after normalization for its respective intrinsic fluorescence as in (C). The observed axial intensity decay (exponential fit, Dz = 481±166 nm) is the result of the increasing defocus for surface-distant beads and the objective's finite depth of field. (D), same, for TIRF excitation (l) and fit of a double-exponential decay (line) with the measured fluorescence F(TIRF)(z; z0) from all beads. Color codes for different θ. The long-range component (Dz) was identical to that measured for out-of-focus beads upon EPI excitation, (D). The short-range component δ(θ) was taken as the effective EW the penetration depth and was, respectively, 349±118, 139±20, 109±16 and 91±13 nm for θ = 65.0°, 67.5, 70.0°, and 72.5°. Note the surface-enhancement by a factor of ~4 of F(TIRF)(z0) vs. F(EPI)(z0). Depending on θ, the fractional amplitude of the non-evanescent long-range excitation component varied between 13 and 17%, see main text. Inset shows double-exponential fits on a log scale and 95% confidence interval of the fit. FIGURE S1. Method for fixing beads on and above a glass coverslip. Prior to the entire procedure a dash with a permanent marker pen was made on the upper face of the cover slip to facilitate focusing at the reflecting interface. Then, left, 1 µl of a 1:5 diluted solution of 0.1-µm diameter TetraSpeckTM was pipetted to the glass coverslip. Beads were left to dry so that they adhered to the glass. Subsequently, middle, 20 µl of agarose/sucrose solution again containing TetraSpeckTM beads at a 1:4 dilution was applied to the coverslip. The agarose was allowed to cool for polymerization. Right, the whole was topped with ~1.5 ml agarose/sucrose (n3 = 1.375) solution previously kept at 35°C. After 10 min at 4°C the entire solution jellified. The result is a low-density carpet of beads attached to the coverslip, super-seeded with beads at different heights above the coverslip. arXiv - presubmission 28 Oheim et al. (2019) TIRF calibration arXiv - presubmission 29 Oheim et al. (2019) TIRF calibration arXiv - presubmission 30 Oheim et al. (2019) TIRF calibration arXiv - presubmission 31 Oheim et al. (2019) TIRF calibration arXiv - presubmission 32 Oheim et al. (2019) TIRF calibration arXiv - presubmission 33 Oheim et al. (2019) TIRF calibration arXiv - presubmission 34 Oheim et al. (2019) TIRF calibration arXiv - presubmission 35 Oheim et al. (2019) TIRF calibration arXiv - presubmission 36 Oheim et al. (2019) TIRF calibration C (iii) (iv) A B (i) (ii) Oheim et al. Supplementary Figure S1 arXiv - presubmission 37 Oheim et al. (2019) TIRF calibration arXiv - presubmission 38
1612.09550
1
1612
2016-12-30T18:25:42
Soliton-like attractor for blood vessel tip density in angiogenesis
[ "physics.bio-ph", "cond-mat.stat-mech", "nlin.PS" ]
Recently, numerical simulations of a stochastic model have shown that the density of vessel tips in tumor induced angiogenesis adopts a soliton-like profile [Sci. Rep. 6, 31296 (2016)]. In this work, we derive and solve the equations for the soliton collective coordinates that indicate how the soliton adapts its shape and velocity to varying chemotaxis and diffusion. The vessel tip density can be reconstructed from the soliton formulas. While the stochastic model exhibits large fluctuations, we show that the location of the maximum vessel tip density for different replicas follows closely the soliton peak position calculated either by ensemble averages or by solving an alternative deterministic description of the density. The simple soliton collective coordinate equations may also be used to ascertain the response of the vessel network to changes in the parameters and thus to control it.
physics.bio-ph
physics
, Soliton-like attractor for blood vessel tip density in angiogenesis 1G. Mill´an Institute, Fluid Dynamics, Nanoscience and Industrial Mathematics, L. L. Bonilla, M. Carretero, and F. Terragni Universidad Carlos III de Madrid, 28911 Legan´es, Spain (Dated: December 16th, 2016) Recently, numerical simulations of a stochastic model have shown that the density of vessel tips in tumor induced angiogenesis adopts a soliton-like profile [Sci. Rep. 6, 31296 (2016)]. In this work, we derive and solve the equations for the soliton collective coordinates that indicate how the soliton adapts its shape and velocity to varying chemotaxis and diffusion. The vessel tip density can be reconstructed from the soliton formulas. While the stochastic model exhibits large fluctuations, we show that the location of the maximum vessel tip density for different replicas follows closely the soliton peak position calculated either by ensemble averages or by solving an alternative deterministic description of the density. The simple soliton collective coordinate equations may also be used to ascertain the response of the vessel network to changes in the parameters and thus to control it. PACS numbers: 87.19.uj, 87.85.Tu, 05.45.Yv, 05.45.-a I. INTRODUCTION The growth of blood vessels is a complex multiscale process called angiogenesis that is the basis of organ growth and repair in healthy conditions and also of pathological developments such as cancerous tumors [1 -- 4]. Cells in an incipient tumor located in tissue expe- rience lack of oxygen and nutrients, and stimulate pro- duction of vessel endothelial growth factor that, in turn, induces growth of blood vessels (angiogenesis) from a nearby primary vessel in the tumor direction [1, 2]. Blood brings oxygen and nutrients that foster tumor growth. In angiogenesis, events happening in cellular and subcellular scales unchain endothelial cell motion and proliferation, build millimeter scale blood sprouts and networks thereof [3, 5 -- 7]. Angiogenesis imbalance contributes to numer- ous malignant, inflammatory, ischaemic, infectious, and immune disorders [2]. For these reasons, immense hu- man and material resources are devoted to understanding and controlling angiogenesis. Theoretical efforts based on angiogenesis models go hand in hand with experiments [8 -- 31]. Models range from very simple to extraordinar- ily complex and often try to illuminate some particular mechanism; see the review [31]. Realistic microscopic models involve postulating mechanisms and a large num- ber of parameters that cannot be directly estimated from experiments, but they often yield qualitative predictions that can be tested. An important challenge is to extract mesoscopic and macroscopic descriptions of angiogenesis from the diverse microscopic models. Early angiogenesis macroscopic models consisted of reaction-diffusion equations for densities of cell and chemicals (growth factors, fibronectin, etc.) [8, 10, 11]. These models do not allow to treat the growth and evo- lution of individual blood vessels. Later models focused on the evolution of the cells at the tip of a vessel sprout. The ten or so cells at a vessel tip are highly motile and do not proliferate. They follow chemotactic and haptotactic clues as they advance toward hypoxic regions that expe- rience lack of oxygen. These cells are followed by prolifer- ating stalk cells that build a capillary in their wake. Thus tip cell models are based on the motion of single particles representing the tip cells and their trajectories constitute the advancing blood vessels [9, 12, 15, 16, 21, 30 -- 32]. More realistic and necessarily more complex models il- luminate tip and stalk cell dynamics, the motion of tip and stalk cells on the extracellular matrix outside blood vessels, blood circulation in newly formed vessels, and so on [20, 22, 29, 31]. In recent work [30, 32], we have been trying to bridge the gap between microscopic descriptions of early stage tumor induced angiogenesis that require large numeri- cal simulations and macroscopic descriptions that are amenable to a more thorough theoretical study. We consider a simple tip cell model in which tip stochas- tic extension is driven by the gradient of growth factors (chemotaxis), there is a random branching of tips and tips join with existing blood vessels (anastomosis). We have derived a deterministic description for the density of vessel tips consisting of an integrodifferential equa- tion for the tip density coupled to a reaction-diffusion equation for the tumor angiogenic factor (TAF, which comprises vessel endothelial and other growth factors) [30, 32]. The stochastic model can be made more real- istic by adding equations characterizing haptotaxis, the influence of other chemicals or drugs, etc. While cell densities can be extracted from numerical simulations of microscopic models, our equation for the tip density [30] incorporates tip branching and anastomosis as derived from a stochastic model [32], not postulated ad hoc. It turns out that the tip density soon forms a moving lump that advances towards the tumor. The longitudinal sec- tion of the stable lump (that we may term angiton) is approximately given by a moving soliton-like wave [33]. This wave is an exact 1D solution of a reduced equation for the marginal tip density on the whole real line that has constant chemotactic force and no diffusion. It ap- pears by differentiating a domain-wall solution (topologi- cal soliton) connecting two spatially homogeneous states. Numerical evidence shows that it is asymptotically stable [33]. Technically speaking, it is not known whether two soliton-like waves in the angiogenesis model equations emerge unchanged from collisions except for a phase shift. Therefore we do not claim that angiogenesis soliton-like lump profiles are true solitons. However stable soliton- like waves are central to the arguments of the present paper and, by an abuse of language, we will call them solitons. In this, we follow extended usage in the physi- cal literature in which other stable waves such as "topo- logical solitons" [34] or "diffusive solitons" [35] are called simply solitons despite not emerging unscathed from col- lisions [34, 35]. The soliton shape and velocity depend on two collective coordinates. The vessel tip density ap- proaches the soliton solution after an initial formation stage. After its formation and until the vessels are close to the tumor, the tip density is described by the soliton and the solution of its two collective coordinate equa- tions. In this paper, we deduce the equations for the angio- genesis soliton and its collective coordinates, solve the latter numerically and reconstruct the marginal tip den- sity from the soliton formula. Then we show that it agrees with both the solution of the deterministic de- scription and with the ensemble average of the tip density as extracted from the stochastic process. Although the fluctuations are large, we give numerical evidence that the position of the soliton peak is very close to that of the maximum of the marginal tip density for different replicas or realizations of the stochastic process. This implies that the simple description based on the soliton may give useful information about single replicas of the angiogenesis process. While our simple model needs to be completed to discuss control of angiogenesis, we show how changing a single parameter results in seemingly ar- resting the process. The rest of the paper is as follows. We recall the stochastic model of [30] and its deterministic description [32] in Section II. By a Chapman-Enskog method, we derive a reduced equation for the marginal tip density in Section III. By neglecting diffusion and considering constant coefficients in the resulting equation, we find in Section IV an analytical expression for the soliton of the marginal tip density [33]. Section V contains a derivation of the differential equations for the two collective coordi- nates of the soliton. The coefficients appearing in these equations contain spatial averages of the TAF density. In Section VI, we explain how to calculate the coeffi- cients in the collective coordinate equations, solve them numerically, reconstruct the soliton and, through it, the marginal vessel tip density. We compare it with direct solutions of the deterministic description and ensemble averages of the stochastic process. Although realizations of the stochastic angiogenic process provide very different looking vessel networks, we also show that the maximum of the marginal density for each realization follows closely 2 the soliton peak. Section VII contains our conclusions and the Appendices are devoted to technical matters. II. MODEL Early stages of angiogenesis are described by a sim- ple stochastic model in [30, 32]. It consists of a system of Langevin equations for the extension of vessel tips, a tip branching process and tip annihilation (anastomosis) when they merge with existing vessels. A tip i is born at a random time T i from a moving tip (we ignore branching from mature vessels) and disappears at a later random time Θi, either by reaching the tumor or by anastomo- sis. At time T i, the velocity of the newly created tip i is selected out of a normal distribution, δσv (v − v0) = e−v−v02/σ2 v πσ2 v , (1) with mean v0 and a narrow variance σ2 v. In addition, the probability that a tip branches from one of the ex- isting ones during an infinitesimal time interval (t, t + dt] i=1 α(C(t, Xi(t)))dt, where is taken proportional to PN (t) C(t, x) is the TAF concentration and α(C) = α1 , CR > 0, α1 > 0, (2) C CR + C in which CR is a reference concentration. The change per unit time of the number of tips in boxes dx and dv about x and v is N (t) Xi=1 α(C(t, Xi(t))) δσv (vi(t) − v0) =ZdxZdv α(C(t, x)) N (t) ×δσv (v − v0) δ(x − Xi(t))δ(v − vi(t))dxdv. (3) Xi=1 The Langevin equations for tip extensions are dXi(t) = vi(t) dt, dvi(t) =(cid:2)−k vi(t) + F(cid:0)C(t, Xi(t))(cid:1)(cid:3)dt + σ dWi(t),(4) where Xi(t) and vi(t) are the tip position and veloc- ity of tip i at time t, Wi(t) are independent identically distributed (i.i.d.) standard Brownian motions, and k (friction coefficient) and σ are positive parameters. At each time t there are N (t) active tips. The chemotactic force is F(C) = d1 (1 + γ1C)q ∇xC, (5) where d1, γ1, and q are positive parameters. The TAF concentration solves ∂ ∂t C(t, x) = d2∆xC(t, x) − ηC(t, x) N (t) Xi=1 × (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) vi(t)δσx (x − Xi(t))(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) . (6) pN(t, x, v) = 1 N δσx(x − Xi(t, ω)) × δσv (v − vi(t, ω)), Xω=1 Xi=1 N N (t,ω) N Xω=1 Xω=1 Xi=1 Xi=1 N (t,ω) pN (t, x) = jN (t, x) = 1 N 1 N δσx(x − Xi(t, ω)), (7) (8) vi(t, ω)δσx(x − Xi(t, ω)).(9) ∂ ∂t p(t, x, v) = A C(t, x) 1 + C(t, x) p(t, x, v)δv(v − v0) 0 Z p(s, x, v′) dv′ds − v · ∇xp(t, x, v) [1 + Γ1C(t, x)]q − βv(cid:19) p(t, x, v)(cid:21) −Γp(t, x, v)Z t −∇v ·(cid:20)(cid:18) δ ∇xC(t, x) + ∆vp(t, x, v), β 2 ∂ ∂t C(t, x) = κ∆xC(t, x) − χ C(t, x)j(t, x), (14) (15) respectively. The dimensionless parameters are defined in Table II and the boundary conditions to solve (14)- (15) are listed in Appendix A. Here d2 (diffusivity) and η are positive parameters, whereas δσx(x) is a regularized smooth delta function (e.g., a Gaussian with variances l2 y proportional to σ2 x along the x and y directions, respectively) that becomes δ(x) in the limit as σx → 0. x and l2 There is a counterpart to the stochastic model for the densities of vessel tips and the vessel tip flux, defined as ensemble averages over a sufficient number N of replicas (realizations) ω of the stochastic process: 3 p 1 L2 j v0 L2 C p x v v0 t L v0 CR 1 L v2 0 L2 mm µm/hr hr mol/m2 1021 s2 2.025 2 50 10−16 40 m4 105m−2 m−1s−1 0.0028 2.5 TABLE I: Units for nondimensionalizing the model equations. N N (t,ω) tion as in Table I [30, 32], (10) and (12) become As N → ∞, these ensemble averages tend to the tip density p(t, x, v), the marginal tip density p(t, x), and the tip flux j(t, x), respectively. In [32] it is shown that the angiogenesis model has a deterministic description based on the following equation for the density of vessel tips, p(t, x, v), p(t, x, v) = α(C(t, x)) p(t, x, v)δv (v − v0) δ ∂ ∂t −γ p(t, x, v)Z t 0 p(s, x) ds − v · ∇xp(t, x, v) −∇v · [(F(C(t, x)) − kv)p(t, x, v)] d1CR β kL v0 A α1 L v3 0 Γ γ v2 0 v2 0 1.5 5.88 22.42 0.145 Γ1 γ1CR κ d2 v0L χ η L σv - 1 0.0045 0.002 0.08 + σ2 2 ∆vp(t, x, v), p(t, x) =Z p(t, x, v′) dv′. The TAF equation (6) becomes (10) (11) ∂ ∂t C(t, x) = d2∆xC(t, x) − η C(t, x)j(t, x), (12) where j(t, x) is the current density (flux) vector at any point x and any time t ≥ 0, j(t, x) =Z v′p(t, x, v′) dv′. (13) Alternatively, if N (t) becomes very large (which is pre- cluded by anastomosis), the same deterministic descrip- tion can be derived by using the law of large numbers [30]. The deterministic description consisting of Equations (10) and (12) is well posed, as it has been proved to have unique smooth solutions [36]. After nondimensionaliza- TABLE II: Dimensionless parameters. III. REDUCED EQUATION FOR THE MARGINAL TIP DENSITY We can obtain a simpler equation for the marginal ves- sel tip density (11) provided the overall tip density ap- proaches rapidly a local equilibrium which is a displaced Maxwellian: p(0)(t, x, v) = 1 π e−v−v02 p(t, x). (16) The source terms in (14) (two first terms on its right hand side) select velocities on a small neighborhood of v0, as such velocities are the only ones for which the birth term proportional to α(C)δv(v − v0), cf Eq. (1), can compensate the anastomosis death term. To derive the simpler equation for p, we use the Chapman-Enskog method [37]. We first rewrite (14) as Lp ≡ β ∇v ·(cid:18) 1 2 ∇vp + (v − v0)p(cid:19) + β (F − v0) · ∇vp + v∇xp ∂t = ǫ(cid:20) ∂p − αp δv(v − v0) + ΓpZ t 0 α = F = , A C 1 + C δ β ∇xC(t, x) [1 + Γ1C(t, x)]q . p(s, x) ds(cid:21), (17) (18) (19) We have included a scaling parameter ǫ in the right hand side of (17), as we will consider that it is small compared to the left hand side. After the computations that follow, we will restore ǫ = 1. Note that (16) satisfies Lp(0) = 0, (20) i.e., (17) with ǫ = 0. We now assume that the terms on the right hand side of (17) are small compared to those on its left hand side (formally, ǫ ≪ 1) and that we can expand p in the asymptotic series 4 as the adjoint problem L†v = 0 has constant solutions. For (24), this condition yields F (0) = α π p − v0 · ∇x p − ΓpZ t 0 p(s, x) ds, (27) which, inserted back in (24), produces the equation Lp(1) = e−V 2 π (cid:26)α(cid:20) 1 π − δv(V)(cid:21)p + V ·[∇x p − 2β(F − v0)p]} . The solution of (28) that satisfies (22) is e−V 2 π p(1) = − V·[∇x p − 2β(F − v0)p] + αp 2π2 e−V 2(cid:20)Z ∞ 0 e−t ln t dt − ln V 2(cid:21). (28) (29) Insertion of (29) into the solvability condition (26) for j = 2 produces F (1) = 1 2β ∆x p + ∇x ·[(v0 − F) p] + α2 p 2π2β(1 + σ2 v) ln(cid:18)1 + 1 σ2 v(cid:19). (30) p = p(0) + ǫp(1) + ǫ2p(2) + . . . . (21) We now substitute (27) and (30) in (23) and recall ǫ = 1, thereby finding the Smoluchowski-type equation Inserting this into (11), we find Z p(j)dv = 0, We assume now that j = 1, 2, . . . . (22) ∂ p ∂t = F (0) + ǫF (1) + . . . , (23) where the F (j) should be determined by solvability con- ditions to be derived below. Inserting (21) and (23) in (17) and equating like powers of ǫ in the result, we obtain the hierarchy of equations (20) and e−V 2 Lp(1) = π hF (0) + v · ∇x p − 2βV·(F − v0)p −αpδv(V) + ΓpZ t 0 p(s, x) ds(cid:21), Lp(2) = F (1) + v · ∇xp(1) e−V 2 π −2βV·(F − v0)p(1) − αp(1)δv(V) +Γp(1)Z t 0 p(s, x) ds, (24) (25) etc. Here V = v − v0 and V = V. For these equa- tions to have bounded solutions, we need to impose the conditions ∂ p ∂t + ∇x · (Fp) − 1 2β ∆x p = µ p −ΓpZ t 0 p(s, x) ds, (31) µ = α π (cid:20)1 + α 2πβ(1 + σ2 v) ln(cid:18)1 + 1 σ2 v(cid:19)(cid:21). (32) Note that the convective terms in (31) correspond to hav- ing ignored inertia in the Langevin equation (4), which then becomes dXi(t) = (F/k) dt + (σ/k) dWi(t). Our perturbation procedure just renormalizes the birth term α(C) in (14) or (17). The flux (13) in the reaction-diffusion equation (15) is j(t, x) ≈ v0 p(t, x), so that (15) becomes ∂ ∂t C(t, x) = κ∆xC(t, x) − χ C(t, x) p(t, x), (33) because v0 = 1 in our nondimensional units. The boundary conditions for (31) are: (i) p(t, x) known at x = 1 and equal to its instantaneous value there; and (ii) known flux j0 at x = 0 [30]. The boundary condi- tion (i) is a free boundary condition that avoids model- ing explicitly the tumor instead of the more appropriate absorbing boundary condition p = 0 at the tumor. In condition (ii), the flux can be approximated as Z (v0 + V)p(t, x, v) dV = v0 p +Z Vp(1)dV 1 2β ∇x p. Z Lp(j)dv = 0, j = 1, 2, . . . , (26) = Fp − At x = 0, the x-component of F is zero and therefore the boundary condition for p becomes − 1 2β ∂ p ∂x = j0, i.e., Here ξ0 is a constant of integration. Thus p = ∂ρ yields ∂t = −c ∂ρ ∂ξ 5 − 1 2β ∂ p ∂x(cid:12)(cid:12)(cid:12)(cid:12)x=0 = v0µ p θ(τ − t), (34) in which θ(t) = 1 if t > 0 and θ(t) = 0 otherwise is the unit step function. In (34), we have renormalized the birth rate coefficient α to µ in harmony with the change in birth rate when going from the equation for the vessel tip density (14) to (31) for the marginal vessel tip density; see (A.6) in Appendix A. In [30, 32] and in the numerical calculations of this paper, τ = ∞. IV. SOLITON We now find an approximate soliton solution of (31) following [33]. Firstly, let define p = (2KΓ + µ2)c 2Γ(c − Fx) sech2"p2KΓ + µ2 2(c − Fx) (x − ct − ξ0)#.(42) This is similar to the usual soliton solution of the Korteweg-de Vries equation except that we now have three parameters, c, K and ξ0. Note that the soliton appears as consequence of a dominant balance of time derivative, convection, and source terms in (31). The existence of the soliton solution is consequence of the quadratic anastomosis term in (14) first derived in [30]. While simulations of the deterministic [30] and stochas- tic descriptions [32] clearly exhibit a soliton-like solution, the derivation presented here first appeared in [33]. ρ(t, x) =Z t 0 p(s, x) ds, (35) V. COLLECTIVE COORDINATES and ignore diffusion in (31), which then becomes ∂2ρ ∂t2 + ∇x ·(cid:18)F ∂ρ ∂t(cid:19) = µ ∂ρ ∂t − Γρ ∂ρ ∂t . (36) The coefficients κ and χ in (33) are very small [30] and therefore the TAF concentration varies very slowly com- pared with the marginal tip density. We will also assume that the initial TAF concentration varies on a larger spa- tial scale than the soliton size and that the TAF gradient is directed on the x axis, which constitutes a good ap- proximation [30]. Then F and µ are almost constant and we will seek a solution of the form ρ(t, x) = ρ(ξ), ξ = x − ct, (37) for (36). The resulting ordinary differential equation is (c − Fx) ∂2ρ ∂ξ2 + (µ − Γρ) ∂ρ ∂ξ = 0, (38) in which Fx is the x-component of the chemotactic force F. Integrating (38) once, we obtain (c − Fx) ∂ρ ∂ξ +(cid:18)µ − Γ 2 ρ(cid:19)ρ = −K, where K is a constant. From this, we get (c − Fx) 2 Γ ∂ρ ∂ξ = ρ2 − 2 µ Γ ρ − 2K Γ . Setting ρ = µ 2νλ(c − Fx)/Γ = −ν2, thereby obtaining Γ + ν tanh(λξ), we find ν2 = µ2+2KΓ Γ2 (39) (40) and ρ = µ Γ − p2KΓ + µ2 Γ tanh"p2KΓ + µ2 2(c − Fx) (ξ − ξ0)#. (41) In this section, we shall discuss the effect of small dif- fusion and a slowly varying TAF concentration on the soliton. Let the soliton solution (42) be written as ps = (2KΓ + µ2)c 2Γ(c − Fx) sech2s, (43) s = p2KΓ + µ2 2(c − Fx) dX dt = c. X = ξ, ξ = x − X(t), (44) (45) Here X(t), c(t) and K(t) are time-dependent collective coordinates characterizing the soliton. They are sup- posed to vary slowly so that the marginal tip density is described by a soliton that moves and changes shape slowly according to the changes of its collective coordi- nates. To find equations for them, we adapt the per- turbation method explained in References [38, 39]. Note that ps is a function of ξ and also of x and t through C(t, x), ps = ps(cid:18)ξ; K, c, µ(C), Fx(cid:18)C, ∂C ∂x(cid:19)(cid:19). (46) We assume that the time and space variations of C, which appear when ps is differentiated with respect to t or x, produce terms that are small compared to ∂ ps/∂ξ. As indicated in Appendix B, we shall consider that µ(C) is approximately constant, ignore ∂C/∂t because the TAF concentration is varying slowly (the dimensionless coeffi- cients κ and χ appearing in the TAF equation (33) are very small according to Table II) and ignore ∂2 ps/∂i∂j, where i, j = K, Fx. Appendix C explains what happens if we relax these assumptions. We now insert (43) and (44) into (31), thereby obtaining + ∂ ps ∂K K + ∂ ps ∂c c + ps∇x · F ∂ξ (cid:16)Fx − X(cid:17) ∂ ps ∂Fx(cid:18) ∂Fx ∂ ps ∂t + + F · ∇xFx(cid:19)− 1 ∂ξ2 2β(cid:18) ∂2 ps ∆xFx(cid:19) = µps +2 ∂2 ps ∂ξ∂Fx ∂Fx ∂x + ∂ ps ∂Fx −ΓpsZ t 0 psdt. (47) Eq. (31) with 1/β = 0 and constant F has the soliton solution (43)-(44). Using this fact and (45), (47) becomes ∂ ps ∂K A = + K + c = A, ∂ ps ∂c ∂2 ps ∂ξ2 − ps∇x ·F− ∂2 ps ∂ξ∂Fx ∂Fx ∂x . 1 2β 1 β (48) ∂ ps ∂Fx(cid:20)F·∇xFx − 1 2β ∆xFx(cid:21) (49) See Appendix B for the precise meaning of these equa- tions. We now find collective coordinate equations (CCEs) for K and c. As the lump-like angiton moves on the x axis, we set y = 0 to capture the location of its maxi- mum. On the x axis, the profile of the angiton is the soliton (43)-(44). We first multiply (48) by ∂ ps/∂K and 6 integrate over x. We consider a fully formed soliton far from primary vessel and tumor. As it decays exponen- tially for ξ ≫ 1, the soliton is considered to be localized on some finite interval (−L/2, L/2). The coefficients in the soliton formulas (43)-(44) and the coefficients in (48) depend on the TAF concentration at y = 0, therefore they are functions of x and time and get integrated over x. The TAF varies slowly on the support of the soliton, and therefore we can approximate the integrals over x by ZI F (ps(ξ; x, t), x)dx ≈ 1 LZI Z L/2 −L/2 F (ps(ξ; x, t), x)dξ!dx. (50) See Appendix B. The interval I over which we integrate should be large enough to contain most of the soliton, of extension L. Thus the CCEs hold only after the initial soliton formation stage. Near the tumor, the boundary condition affects the soliton and we should exclude an interval near x = 1 from I. We shall specify the inte- gration interval I in the next section. Acting similarly, we multiply (48) by ∂ ps/∂c and integrate over x. From the two resulting formulas, we then find K and c as frac- tions. The factors 1/L cancel out from their numerators and denominators. As the soliton tails decay exponen- tially to zero, we can set L → ∞ and obtain the following CCEs [33] K = R ∞ −∞ c = R ∞ −∞ −∞ ∂ ps ∂c (cid:17)2 −∞(cid:16) ∂ ps ∂K AdξR ∞ dξ−R ∞ ∂K(cid:17)2 ∂c (cid:17)2 −∞(cid:16) ∂ ps −∞(cid:16) ∂ ps R ∞ dξR ∞ ∂K(cid:17)2 −∞(cid:16) ∂ ps ∂c AdξR ∞ dξ−R ∞ ∂c (cid:17)2 ∂K(cid:17)2 −∞(cid:16) ∂ ps −∞(cid:16) ∂ ps R ∞ dξR ∞ ∂ ps −∞ ∂ ps ∂K ∂ ps ∂c dξ , ∂ ps ∂c AdξR ∞ −∞ ∂ ps ∂c dξ−(cid:16)R ∞ −∞ ∂ ps ∂K dξ(cid:17)2 ∂ ps ∂K ∂ ps ∂c dξ . ∂ ps ∂K AdξR ∞ −∞ ∂ ps ∂c dξ−(cid:16)R ∞ −∞ ∂ ps ∂K dξ(cid:17)2 In these equations, all terms varying slowly in space have been averaged over the interval I. The last term in (49) is odd in ξ and does not contribute to the integrals in (51) and (52) whereas all other terms in (49) are even in ξ and do contribute. The integrals appearing in (51) and (52) are calculated in Appendix D. The resulting CCEs are c = − 7(2KΓ + µ2) 20β(c − Fx) F·∇xFx − (c − Fx)∇x ·F − ∆xFx 2β (cid:0)1 − 4π2 15 (cid:1)(cid:0)1 − Fx 2c(cid:1) 1 − 4π2 105 + , 2 − Fx c (51) (52) (54) K = (2KΓ+µ2)2 4Γβ(c−Fx)2 − 2KΓ + µ2 2Γc(cid:0)1 − Fx c 75 − 9 4π2 75 + 1 5 +(cid:16) 2Fx (cid:0)1 − 4π2 10(cid:17)Fx 5c − 2π2 2c(cid:1)2 15 (cid:1)(cid:0)1 − Fx 2c(cid:1)(cid:18)c∇x · F+F·∇xFx − 2β (cid:19),(53) ∆xFx in which the functions of C(t, x, y) have been averaged over the interval I and we have set y = 0. We expect the CCEs (53)-(54) to describe the mean behavior of the soliton whenever it is far from primary vessel and tumor. We back this point of view by the numerical simulations reported in the next section. K(t) from deterministic model K(t) from stochastic model (a) 240 200 160 K 120 15 t (hr) 20 c(t) from deterministic model c(t) from stochastic model (b) 15 t (hr) 20 X(t) from deterministic model X(t) from stochastic model (c) 7 25 25 80 40 0 10 2.2 2 1.8 1.6 1.4 1.2 1 0.8 10 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 10 c X VI. NUMERICAL RESULTS Based on numerical simulations [33], we expect that the vessel tip density approaches the soliton after some time. Initially there are few tips and the density is small so that the nonlinear anastomosis terms in (14) or in (31) are small. Tips proliferate and the anastomosis terms kick in. The soliton formation should be described as the solution of a semi-infinite initial-boundary value problem. Ideally, we would match the solution of the soliton for- mation stage with a stage of a soliton moving far from boundaries, which is the crucial stage described by Equa- tions (53)-(54) for the collective coordinates. We expect the soliton solution to be an asymptotically stable solu- tion of the vessel tip density equation (14) on the whole 1D real line and also for the 2D slab geometry considered in this paper (provided the primary vessel is at x = −∞ and the tumor is at x = +∞). For a slowly varying TAF density, the stable soliton will instantaneously adapt its shape and velocity according to the solution of the CCEs (53)-(54). In this paper, we will solve numerically the full equa- tions (14) (with q = 1) and (15) for the vessel tip density and the TAF density (deterministic description), which we will also obtain by ensemble averages from stochastic simulations as explained in [32]. From these simulations, we will obtain the evolution of the soliton collective coor- dinates thereby reconstructing the marginal tip density at y = 0 from (43). The soliton provides a simple de- scription of tumor induced angiogenesis that agrees with numerical simulations of the stochastic process and with numerical simulations of the deterministic description. Both deterministic or stochastic simulations show that the soliton is formed after some time t0 = 0.2 (10 hours) following angiogenesis initiation. To find the soliton evo- lution afterwards, we need to solve the CCEs (53)-(54) whose coefficients are spatial averages over a certain in- terval x ∈ I that depend on the TAF concentration C(t, x, y) and its derivatives calculated at y = 0. The interval I should exclude regions affected by boundaries. We calculate the spatially averaged coefficients in (53)- (54) by: (i) approximating all differentials by second or- der finite differences, (ii) setting y = 0, and (iii) aver- aging the coefficients from x = 0 to 0.6 by taking the arithmetic mean of their values at all grid points in the interval I = (0, 0.6]. For x > 0.6, the boundary condi- tion at x = 1 influences the outcome and therefore we leave values for x > 0.6 out of the averaging. The initial conditions for the CCEs (45), (53) and (54) are set as follows. X(t0) = X0 is the location of the marginal tip density maximum, p(t0, x = X0, 0). We find X0 = 0.22 from the deterministic description and X0 = 0.2 from the stochastic description. We set c(t0) = c0 = X0/t0. K(t0) = K0 is determined so that the maximum marginal tip density at t = t0 coincides with the soliton peak. This yields K0 = 173 (deterministic description) and 39 (stochastic description). Solving the CCEs (45), (53) and (54) with these initial conditions, we obtain the 15 t (hr) 20 25 FIG. 1: Evolution of the collective coordinates: (a) K(t), (b) c(t), and (c) X(t). curves depicted in Figure 1. Using the soliton collective coordinates depicted in Fig- ure 1 and (43)-(44), we reconstruct the marginal vessel tip density and find its maximum value and the location thereof for all times t > t0. Figure 2 shows that the soli- ton as predicted from the CCEs (45), (53) and (54) com- pares very well with the tip density obtained by direct numerical simulation of the deterministic equations. An alternative way to find the coefficients of the CCEs and their proper initial conditions is to use ensemble averages of the stochastic process. Figure 3 shows that such recon- struction of the soliton agrees very well with the vessel tip density provided by ensemble averages of the stochas- tic process during the 14 hour time interval when soliton motion is not affected by boundaries. There is a large discrepancy between the maximum marginal tip density as predicted by the soliton and by the stochastic process during the first 10 hours of angiogenesis, which clearly marks the duration of the initial stage of soliton forma- tion. After this stage, we note that the location of the maximum of the marginal tip density is very closely pre- maximum of p(t, x, 0) soliton peak 430 410 390 (a) 1100 1000 900 800 700 600 500 400 300 8 maximum of p(t, x, 0) soliton peak 460 440 420 400 380 10 12 14 20 22 24 16 18 t (hr) (a) 370 10 12 14 16 18 t (hr) 20 22 24 200 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34 36 t (hr) position of p(t, x, 0) peak soliton peak position 0.7 0.6 0.5 0.4 0.3 position of p(t, x, 0) peak soliton peak position 0.6 0.5 0.4 0.3 (b) (b) 0.2 10 12 14 16 18 t (hr) 20 22 24 0.2 10 12 14 16 18 t (hr) 20 22 24 FIG. 2: Deterministic description: Comparison between the maximum value of p(t, x, 0) and its value as predicted by soli- ton collective coordinates. (a) Evolution of the maximum value of the marginal tip density (relative error smaller than 4.5%). (b) Evolution of the position of the maximum marginal tip density on [0, 1] (at t = 20 and 22 h, the absolute error is the space step in the numerical method, ∆x = 0.02; at t = 24 h, the error is 4∆x). FIG. 3: Same as in Figure 2 for the stochastic description. The zoom in Figure 3(a) corresponds to Figure 2(a) but we have drawn the same figure with a larger time span to show more clearly the time interval over which the soliton approx- imates the maximum marginal tip density. The relative error is smaller than 6.7% for the maximum marginal tip density (calculated by ensemble average over 400 realizations [32]), whereas the error in the predicted position of the maximum marginal tip density is ∆x = 0.02 at 22h and 2∆x at 24 h. dicted by the location of the soliton peak as a function of time, both by using ensemble averages of the stochastic process as in Figure 3 or by solving numerically the de- terministic description as in Figure 2. This is also clearly shown in the reconstruction of the soliton marginal tip density depicted in Figure 4. So far, our reconstructions have been based on ensem- ble averages or, what is quite similar, the marginal tip density as given by the deterministic description. In past work [32], we have shown that fluctuations about the mean are large and therefore the stochastic process is not self-averaging for a single realization: anastomosis precludes the formation of a large number of active tips that may enforce mean-field behavior. However a deter- ministic description is still possible for averages over a sufficiently large number of realizations of the stochastic process (four hundred realizations suffice), as explained extensively in [32]. This raises an important question: How well do these ensemble averages and the soliton con- struction represent single replicas of the stochastic pro- cess? Figure 5 gives a positive answer for the location of the soliton peak: The position of the soliton peak is a good approximation to the location of the maximum marginal tip density for different replicas of the stochas- tic process. While vessel networks may differ widely from replica to replica, the position of the maximum marginal tip density is about the same for different replicas. As the maximum of the marginal tip density is a good measure of the advancing vessel network, the soliton peak location also characterizes it. The existence of other seemingly self-averaging quantities related to the soliton is an open (a) 36 30 24 18 ) r h ( t 12 6 0 0 36 30 24 18 12 6 0 0 ) r h ( t 0.5 x/L 1 (c) 0.5 x/L 1 36 30 24 18 12 6 0 0 36 30 24 18 12 6 0 0 0.5 x/L 1 FIG. 4: Comparison of the marginal tip density profile to that of the moving soliton for (a) and (b): Deterministic descrip- tion; (c) and (d): Stochastic description averaged over 400 replicas. question. We can use the soliton construction as a simple means to evaluate the influence of new mechanisms on angio- genesis. For instance, suppose that some drug causes the friction coefficient β to increase fivefold. Then the marginal tip density gets delayed as shown by Figure 6. This can be evaluated easily and cheaply by solving the CCEs. What does this mean for replicas of the angio- genesis process? Figure 7 displays the vessel networks formed after 36 hours for β = 5.88 and 29.4 in two dif- ferent replicas of the stochastic process. For β = 5.88, the vessel network of one replica of the angiogenesis pro- cess has reached the tumor at x = 1 after 36 hours, for β = 29.4 the vessel network is only half way through its road to the tumor after that time. Had the increase in β been the result of some therapy, we could have ascer- (b) soliton 0.7 0.6 0.5 0.4 0.3 0.2 0.1 10 9 position of p(t, x, 0) peak (replica 10) position of p(t, x, 0) peak (replica 19) position of p(t, x, 0) peak (replica 100) position of p(t, x, 0) peak (replica 350) soliton peak position 12 14 16 18 t (hr) 20 22 24 0.5 x/L 1 (d) soliton FIG. 5: Position of the soliton peak density compared to that of the maximum marginal tip density for different replicas of the stochastic process. β = 5.88, t = 24 hr β = 5.88, t = 36 hr β = 29.4, t = 24 hr β = 29.4, t = 36 hr 500 400 300 200 100 0 0 0.2 0.4 0.6 0.8 1 x/L FIG. 6: Marginal vessel tip density profiles at 24 (dashed lines) and 36 hours (solid lines) for β = 5.88 (blue lines), and β = 29.4 (red lines). tained its merits by solving the CCEs and inferring the arrest of the vessel network from the result. VII. CONCLUSIONS Previous work has shown that a simple stochastic model of tumor induced angiogenesis could be described deterministically by an integrodifferential equation of Fokker-Planck type with a linear birth term and a non- linear death (anastomosis) term [30, 32]. Anastomosis keeps the number of vessel tips rather small (about one hundred) and therefore the vessel tip density has to be reconstructed from ensemble averages of the stochastic process, which is not self-averaging. Numerical simula- tions of stochastic and deterministic equations show that 0.5 L / y 0 −0.5 0 (a) (b) 0.5 L / y 0 0.5 x/L −0.5 0 1 0.5 x/L 1 FIG. 7: Comparison between the vessel networks of two repli- cas after 36 hours for (a) β = 5.88, and (b) β = 29.4. The TAF level curves have also been depicted. the vessel tip density advances from the primary vessel towards the tumor as a stable moving lump or angiton whose profile along the x axis is soliton-like [30, 32, 33]. An analytic formula for the longitudinal profile of the an- giton (called the "soliton" in this paper) can be deduced by ignoring spatio-temporal variation of the tumor angio- genic factor and diffusion [33]. This formula involves two collective coordinates that characterize the shape and ve- locity of the soliton [33]. In the present work, we have derived the reduced equa- tion for the marginal tip density by means of a Chapman- Enskog method. We have deduced the differential equa- tions for the collective coordinates whose terms involve spatial averages over the fully grown soliton far from the tumor. We can deduce these equations both from the de- terministic description and from ensemble averages of the full stochastic model. In both cases, the soliton provides a good reconstruction of the deterministic marginal tip density or its version based on ensemble averages, pro- vided the soliton is not too close to the tumor. As said before, fluctuations are large because anastomosis keeps a small number of active vessel tips at all time. Never- theless, we have shown that the position of the maximum marginal tip density as given by the soliton is quite close to that given by any replica of the stochastic angiogenesis process. This indicates that the simple soliton construc- tion yields good predictions of the evolution of the blood vessel network. There are mechanisms not included in our stochastic conceptual model of angiogenesis. However, many mech- anisms such as haptotaxis can be included by adding terms to the force F in the Langevin equation for the vessel tips that depend on additional continuum fields (fibronectin, matrix degrading enzymes, etc; see e.g., [21]). The effects of anti-angiogenic factors could be treated by including additional reaction-diffusion equa- tions and their effects on the vessel tips [14]. Such terms can be straightforwardly incorporated to the equa- tions for the soliton collective coordinates using the same methodology as explained in the present paper. There 10 are other models that postulate reinforced random walks [12, 15, 16] or cellular Potts models with Monte Carlo dynamics [20, 25, 31] instead of Langevin equations to describe the extension of vessel tips. Insofar as Fokker- Planck equations can be derived from master equations in appropriate limits [40] and branching and anastomosis are similar to those of our conceptual model, we could use the same methodology as in the present paper to study such models. Let us recall that the soliton solu- tion comes through a balance of birth and death terms, convection and time derivative terms in the equation for the marginal tip density. These terms would also appear in special limits of the random walk or cellular Potts models. We consider the work presented in this paper a blueprint for using the soliton methodology to analyze more complex angiogenesis models and a first step to control angiogenesis through soliton dynamics. Acknowledgments We thank Vincenzo Capasso, Bjorn Birnir and Boris Malomed for fruitful discussions. This work has been supported by the Ministerio de Econom´ıa y Competitivi- dad grant MTM2014-56948-C2-2-P. Appendix A: Boundary conditions for the deterministic equations The nondimensional boundary conditions for the TAF are [30] ∂C ∂x (t, 0, y) = 0, ∂C ∂x (t, 1, y) = aL d2CR e−y2L2/b2 (A.1) (b is half the tumor width) and limy→±∞ C = 0. We do not intend to follow the process of angiogenesis beyond the time that vessels tip have arrived at the tumor and therefore we do not give the latter a finite length. We use a Gaussian as the initial condition for the TAF C(0, x, y) = 1.1 e−[(x−1)2L2/c2+y2L2/b2], (A.2) for appropriate b and c. The boundary conditions for the tip density are [30] p+(t, 0, y, v, w) = ×(cid:20)j0(t, y)−Z 0 p−(t, 1, y, v, w) = ×(cid:20)p(t, 1, y)−Z ∞ −∞ p(t, x, v) → 0 as v → ∞, e−v−v02 −∞ −∞ v′e−v ′−v02dv′ dw′ R ∞ 0 R ∞ −∞Z ∞ v′p−(t, 0, y, v′, w′)dv′dw′(cid:21),(A.3) −∞R ∞ R 0 0 Z ∞ p+(t, 1, y, v′, w′)dv′dw′(cid:21), (A.4) −∞ e−v ′−v02dv′ dw′ e−v−v02 (A.5) Appendix B: Derivation of equation (50) and meaning of the CCEs K = where p+ = p for v > 0 and p− = p for v < 0, v = (v, w). An absorbing boundary condition p = 0 on the tumor surface would be more realistic than (A.4). However this would be computationally more costly as we would need to include a slab that extends beyond x = 1. However the difference with the present results would be appreciable at the last stage when the vessel tips arrive at the tumor, something we do not study specifically in the present paper. In (A.3), the tip flux density at x = 0 is [30] j0(t, y) = v0α(C(t, 0, y)) p(t, 0, y, v0, w0) θ(τ − t), (A.6) for the vector velocity v0 = (v0, w0), with v0 = 1. Different from [30], we have included the step function θ(τ − t) in (A.6). With τ = ∞ as in [30, 32], the primary vessel keeps injecting tip density for all time. However, this may be artificial, as the primary vessel does not in- ject any more vessels after t = 0+ in many experiments on early stage angiogenesis. Then τ in (A.6) may be a small time of the order of the time step used in a nu- merical code. The original boundary condition in [30] did not include the unit step function and, as a con- sequence, the deterministic description given by the tip density equation and its boundary conditions had an ar- tificial injection of tip density at x = 0 for all t > 0. The deterministic description including boundary conditions can be proved to have a solution [41]. Let I = (a, b) and let us assume that ξ + X = x/ǫ, with ǫ ≪ 1, for a fixed time. Let us consider the initial value problem dΛ dx = F (p(ξ; x), x), Λ(a) = 0, (B.1) and solve it by using multiple scales x and ξ, and the assumption Λ = Λ(0)(ξ, x) + ǫΛ(1)(x, ξ) + O(ǫ2). We find the hierarchy of equations ∂Λ(0) ∂ξ ∂Λ(1) ∂ξ = 0, = F (p(ξ; x), x) − ∂Λ(0) ∂x , (B.2) (B.3) and so on. (B.2) means that Λ(0) depends only on x. As- suming boundary conditions p(±L/2; x) = 0, (B.3) has a solution bounded in ξ for large ξ provided the integral of its right hand side over ξ is zero, which yields ∂Λ(0) ∂x = 1 LZ L/2 −L/2 F (p(ξ; x), x) dξ. (B.4) Then Λ(0)(b) gives the formula (50), which is typical in homogenization theory. now Eq. Consider = a µ(C(t, x, 0)) dx/(b − a) ≡ µ and a similar defini- tion for Fx. According to the assumptions specified (46) with µ(C) R b 11 ∂C ∂x(cid:19)(cid:19)= ps(ξ; K, c, µ, Fx) (ξ; K, c, µ, Fx)(Fx − Fx)+. . . . (B.5) below (46), we may write ps = ps(cid:18)ξ; K, c, µ, Fx(cid:18)C, ∂ ps ∂Fx + Then ∇x ps = ex ∂ ps ∂ξ (ξ; K, c, µ, Fx) + ∂ ps ∂Fx (ξ; K, c, µ, Fx)∇xFx + . . . , (B.6) and similarly for ∆x ps. Here ex = (1, 0). Using these formulas, we find A in (49) with the following meaning: A = − + 1 2β ∂ ps ∂Fx 1 β ∂2 ps ∂ξ∂Fx ∂2 ps ∂ξ2 (ξ; K, c, µ, Fx) − ps(ξ; K, c, µ, Fx)∇x ·F (ξ; K, c, µ, Fx)(cid:20)F·∇xFx − ∆xFx(cid:21) 1 2β (ξ; K, c, µ, Fx) . (B.7) ∂Fx ∂x Then the CCEs become (2KΓ+µ2)2 4Γβ(c−Fx)2 c 75 − 9 4π2 75 + 1 5c − 2π2 10(cid:17)Fx 5 +(cid:16) 2Fx 2c(cid:17)2 15 (cid:1)(cid:16)1 − Fx (cid:0)1 − 4π2 2 (cid:17)(cid:18)c∇x · F+F·∇xFx − 2β (cid:19),(B.8) ∆xFx − 2KΓ+µ2 2Γ(cid:16)c − Fx 1 − 4π2 105 c = − 7(2KΓ + µ2) 20β(c − Fx) (cid:0)1 − 4π2 15 (cid:1)(cid:16)1 − Fx 2c(cid:17) F·∇xFx − (c − Fx)∇x ·F − ∆xFx 2β + , (B.9) 2 − Fx c as indicated in Section V. Appendix C: Extended collective coordinate equations for a soliton far from primary vessel and tumor In this Appendix, we will find the CCEs without the assumptions that µ is constant and that the time varia- tion of the TAF concentration is negligible. To obtain the CCEs, we need to substitute the soliton (43) into (31). According to (43), the soliton is a function ps = ps(cid:18)ξ; K, c, µ(C), Fx(cid:18)C, ∂C ∂x(cid:19)(cid:19), (C.1) so that we have the expressions: ∇x ps = ex ∆x ps = + ∂ ps ∂ξ ∂2 ps ∂ξ2 + ∂ ps ∂µ ∂ ps ∂K ∇xFx2 + 2 ∂2 ps ∂F 2 x ∂ ps ∂K = + ∂ ps ∂t ∂Fx ∂t K + ∂ ps ∂c ∂C ∂x µC∇xC + ∂ ps ∂Fx µµC ∆xC + (µ2 ∇xFx = ex ∂ ps ∂ξ + ∂ ps ∂K µµC Γ ∇xC + ∂ ps ∂Fx ∇xFx, C + µµCC )∇xC2 Γ + ∆xFx + ∂2 ps ∂K 2 µ2µ2 Γ2 ∇xC2 C ∂ ps ∂Fx ∂C ∂x(cid:19)+ 2 ∂2 ps ∂ξ∂Fx ∂Fx ∂x , µµC Γ (cid:18) ∂2 ps ∂K∂Fx µµC Γ + ∂ ps ∂ξ c − c ∇xC ·∇xFx + ∂ ps ∂K ∂C ∂t + ∂ ps ∂Fx ∂2 ps ∂ξ∂K ∂Fx ∂t , = δ β ∂ ∂t (1 + Γ1C)q = δ β(1 + Γ1C)q(cid:18) ∂2C ∂t∂x − ∂C ∂x ∂C ∂t qΓ1 1 + Γ1C(cid:19) = δ β ∂ ∂x ∂C ∂t (1 + Γ1C)q , in which ex is the unit vector along the x axis and we have used Substituting (48) (with p = ps) in (C.5), we obtain ∂ ps ∂µ = ∂ ps ∂K µ Γ . (C.6) Inserting (C.2)-(C.5) in (31), we obtain 12 (C.2) (C.3) (C.4) (C.5) ∂Fx ∂t = κδ β ∂ ∂x − χδC β(1 + Γ1C)q(cid:18) ∂ ps ∆xC (1 + Γ1C)q − µµC Γ ∂ξ + χδ β ∂ ps ∂K ps ∂ ∂x ∂C ∂x + C (1 + Γ1C)q ∂ ps ∂Fx ∂Fx ∂x (cid:19).(C.8) ∂ξ (cid:16)Fx − X(cid:17) ∂ ps µC(cid:18) ∂C ∂ ps ∂µ ∂t + +F · ∇xFx)− + ∂ ps ∂K K + ∂ ps ∂c c + ps∇x · F + F · ∇xC(cid:19)+ 2β(cid:18) ∂2 ps ∂ξ2 + 2 1 ∂ ps ∂Fx(cid:18) ∂Fx ∂t ∂C ∂x µC ∂2 ps ∂ξ∂µ ∂2 ps ∂F 2 x ∂Fx ∂x + ∂ ps ∂Fx ∆xFx + ∇xFx2 µµC ∆xC + (µµCC + µ2 C )∇xC2 +2 + + ∂2 ps ∂ξ∂Fx ∂ ps ∂K ∂2 ps ∂K 2 Γ µ2µ2 Γ2 ∇xC2 + 2 C = µps − ΓpsZ t 0 psdt. ∂2 ps ∂K∂Fx µµC Γ ∇xC ·∇xFx(cid:19) (C.7) Inserting these equations into (C.7) and using (43) and (44), we obtain (33) with the following A, A = − + + + 1 2β µ2 C + µµCC ∂2 ps ∂ξ2 − ps∇x · F − ∇xC2(cid:21)− ∂2 ps ∂K 2 ∇xC2 + µ2µ2 C 2βΓ2 2βΓ ∂ ps ∂ ps ∂K(cid:20) µµC ∂Fx(cid:20)F · ∇xFx + Γ (cid:18)F · ∇xC + κ∆xC − χC ps − ∂x(cid:18) ∆xC (1 + Γ1C)q(cid:19)− ∂x(cid:18) ∆xC(cid:19) ∆xFx(cid:21) (1 + Γ1C)q(cid:19) ∂ ps 1 2β 1 2β ∇xC ·∇xFx + ∂K∂Fx δχ β κδ β ∂2 ps ∂Fx C ∂ ∂ µµC βΓ ps 1 2β µµC Γβ ∂2 ps ∂F 2 x ∂2 ps ∂ξ∂K ∇xFx2 + δχC β(1 + Γ1C)q ∂Fx 1 β ∂x ∂2 ps ∂ξ∂Fx + ∂C ∂x + ∂Fx ∂Fx(cid:19)2 ∂x (cid:18) ∂ ps + δχC β(1 + Γ1C)q δχµµC C βΓ(1 + Γ1C)q ∂ ps ∂Fx ∂ ps ∂ξ , ∂C ∂x ∂ ps ∂K ∂ ps ∂Fx (C.9) instead of (49). After we calculate the integrals that appear in (51)- (52) as indicated in Appendix D, these equations become: K = (2KΓ + µ2)2 4Γβ(c − Fx)2 + κδ β ∂ ∂x(cid:18) ∆xC 5βΓ2(c − Fx)2(cid:0)1 − 4π2 1 − 2π2 + − ∂Fx ∂x c ∂ C C − µµC µ2µ2 5 + Fx 75 − 9 15 + Fx 4π2 75 + 1 105 − 4Fx C + µµCC − χδc(2KΓ + µ2)2 105 + 8π4 c − Fx (1 + Γ1C)q(cid:19)(cid:21)− c (cid:16) 2Fx 10(cid:17) 5c − 2π2 2c(cid:1)2 (cid:0)1 − 4π2 15 (cid:1)(cid:0)1 − Fx Γ  F·∇xC +(cid:18)κ −  6(1 + Γ1C)q  15 (cid:1)(cid:0)1 − Fx 2c(cid:1)  #+ ∂x(cid:18) (1 + Γ1C)q(cid:19)(cid:20)1 − c (cid:17) C(cid:16) π2  + 2(2KΓ + µ2)(cid:0)1 − Fx 2c(cid:1) 2(c − Fx)(cid:21), 7(2KΓ + µ2)(cid:16)1 − 4π2 105(cid:17) 15 (cid:1)(cid:0)1− Fx 20β(c−Fx)(cid:0)1− 4π2 2c(cid:1) 2(c − Fx)(cid:21) 1 + π2 −(cid:20) µµC∇xC ·∇xFx β(cid:0)1− Fx 2c(cid:1) 15 (cid:16)1 − 4π2 35 (cid:17) 21 (cid:19) × δ(cid:18)1−   ∂x(cid:18) (1 + Γ1C)q(cid:19)(cid:27)− 1− 34π2 105 5(c − Fx) 2(c − Fx)(1 + Γ1C)q F·∇xFx + κδ β 2KΓ + µ2 ∇xFx2 ∇xFx2 1 + 2π2 ∂C ∂x µ2µ2 2π2 − + + + + + + C β 30 ∂ + ∇xC2 2βΓ  µ2 ×(cid:20) µµC∇xC ·∇xFx 2KΓ + µ2 c = − 1 ∂C 5µµC 2Γc(cid:0)1 − Fx 2β(cid:19)∆xC − 1 2β 2KΓ + µ2 ∆xFx ∂xh1 − 2π2 2c(cid:1)(cid:20)c∇x ·F + F·∇xFx − χcC(2KΓ + µ2)h1 − 34π2 17c(cid:1)i  105 (cid:0)1 − 6Fx  3Γ(c − Fx)(cid:0)1 − 4π2 15 (cid:1)(cid:0)1 − Fx 2c(cid:1) 35c i 5 (cid:17)+ 8π2Fx 105(cid:16)17 + 4π2 7 (cid:19)(cid:21)) 3c (cid:18)1 − 90(cid:17)+ F 2 2c (cid:16)1− π2 2c(cid:1)2 (2KΓ + µ2) βΓ(c − Fx)(cid:0)1 − Fx 86π2 315 30 − 3Fx 2KΓ + µ2 2π2 2Fx x 2c2 1− π2 − ∂ ∂x(cid:16) ∆xC (1+Γ1C)q(cid:17)− ∆xFx 2 − Fx c 2β −(c − Fx)∇x ·F 2KΓ + µ2 735 χδc(2KΓ + µ2) 10(1+Γ1C)q(c−Fx)2 2c(cid:1)(cid:26) µµC C 15 (cid:1)(cid:0)1 − Fx 3βΓ(cid:0)1 − 4π2 i 7Ch1+ 2π2(1−4π2) 2c(cid:1)(cid:18)1+ 30(cid:19)∇xC2. ∂Fx ∂x π2 C(c − Fx) 2β(2KΓ + µ2)2(cid:0)1 − Fx 13 (C.10) (C.11) In Table II, the dimensionless coefficients κ and χ ap- pearing in the TAF equation (33) are very small. Then we may ignore terms having these coefficients in the CCEs (C.10)-(C.11), thereby obtaining K = (2KΓ + µ2)2 4Γβ(c − Fx)2 4π2 75 + 1 − µ2 ∆xC 5 + Fx 75 − 9 10(cid:17) 2KΓ + µ2 C + µµCC c (cid:16) 2Fx 5c − 2π2 2 (cid:1)(cid:18)c∇x·F + F·∇xFx − 2c(cid:1)2 2Γ(cid:0)c − Fx 15 (cid:1)(cid:0)1 − Fx (cid:0)1 − 4π2 c (cid:17) C(cid:16) π2 2β (cid:19)+    2Γ(2KΓ + µ2)(cid:0)1 − Fx 2c(cid:1) 90(cid:17)+ F 2 2c(cid:1)2 (2KΓ + µ2)(cid:20) µµC∇xC ·∇xFx 2(c − Fx)(cid:21), 15 + Fx 2KΓ + µ2 ∇xFx2 ∇xC2 µ2µ2 x 2c2 2β − + Γ µµC − Γ (cid:18)F·∇xC − 2c (cid:16)1− π2 βΓ(c − Fx)(cid:0)1 − Fx 30 − 3Fx 1− π2 + c = − 7(2KΓ + µ2) 20β(c − Fx) 1 − 4π2 105 + (cid:0)1 − 4π2 15 (cid:1)(cid:0)1 − Fx 2c(cid:1) µµC∇xC ·∇xFx 2KΓ + µ2 −(cid:20) µ2µ2 C (c − Fx)∇xC2 2(2KΓ + µ2)2 + F·∇xFx − ∆xFx 2β − (c − Fx)∇x ·F 2 − Fx c + ∇xFx2 2(c − Fx)(cid:21) 1 + π2 β(cid:0)1 − Fx 2c(cid:1) 30 . 1 2β ∆xFx(cid:19) (C.12) (C.13) Further simplification leads to (53)-(54). Appendix D: Derivation of the collective coordinate equations 14 The derivatives of ps, given by (43)-(45), which appear in (C.9) are: maximum of p(t, x, 0) soliton peak 415 395 375 355 335 315 10 12 14 16 18 t (hr) 20 22 24 FIG. 8: Same as Figure 2: Comparison between the maximum value of p(t, x, 0) as given by the deterministic description and its value as predicted by soliton collective coordinates that solve (C.10)-(C.11). We can reconstruct the soliton using the extended CCEs (C.10)-(C.11) instead of the simplified CCEs (53)- (54). Somewhat surprisingly, the reconstruction com- pares poorly with the direct solution of the deterministic description. Figure 8 depicts the evolution of the soliton peak when evaluated from (C.10)-(C.11) and the peak of the reduced density p(t, x, 0) as given by the determin- istic description. We observe that the soliton peak de- creases far away from p(t, x, 0). The reason is that K(t) monotonically decreases with t. Instead, K(t) given by (53)-(54) reaches a minimum and it increases as shown in Figure 1(a). Then the soliton peak calculated from the CCEs (53)-(54) and depicted in Figure 2(a) increases after reaching a local minimum and it becomes closer to p(t, x, 0). The discrepancies between the solutions of the different CCEs are caused by the terms propor- tional to µµC in (C.10). In particular, the negative term − µµC to be compensated by any positive term in the equation for K. In turn, the large value of A in Table II ampli- fies the importance of the spatial variation of C, F·∇xC, reflected in that coefficient. In principle, the CCEs are based on the idea that the spatial variations of C, which appear when ps of (43) and (46) is differentiated with respect to x, produce terms that are small compared to ∂ ps/∂ξ. The large value of A contradicts this idea and thus the CCEs (53)-(54) based on setting µC = 0 give better results than (C.10)-(C.11) or (C.12)-(C.13). 2β (cid:17) in (C.10) or in (C.12) is too large Γ (cid:16)F·∇xC − ∆xC = = ∂2 ps ∂ξ2 = ∂ ps ∂K ∂ ps ∂c ∂ ps ∂µ ∂ ps ∂Fx ∂2 ps ∂K 2 = ∂2 ps = = c(2KΓ + µ2)2 4Γ(c − Fx)3 sech4s (2 sinh2 s − 1), (D.1) c (D.2) sech2s (1 − s tanh s), c − Fx c(2KΓ + µ2) Γ(c − Fx)2 sech2s(cid:18)s tanh s − Fx 2c(cid:19), (D.3) (D.4) , ∂ ps ∂K µ Γ c(2KΓ + µ2) Γ(c − Fx)2 sech2s(cid:18) 1 2 − s tanh s(cid:19), (D.5) cΓ s ∂ ∂s [sech2s (1 − s tanh s)] (c − Fx)(2KΓ + µ2) , (D.6) c s (c − Fx)2 ∂ ∂s [sech2s (1 − s tanh s)] = + ∂ ps ∂K , 1 c − Fx 2 ∂ ps ∂Fx + c − Fx ∂ ∂s(cid:20)sech2s(cid:18) 1 2 c(2KΓ + µ2) Γ(c − Fx)3 s − s tanh s(cid:19)(cid:21), ∂K∂Fx ∂2 ps ∂F 2 x = × and (D.7) (D.8) (D.9) µC ≡ ∂µ ∂C = d π(1 + C)2 1 + σ2 πβ(1 + σ2 v(cid:17) α ln(cid:16)1 + 1 v)  . To find the CCEs of the soliton, we need the following integrals calculated from (D.1)-(D.9): ∂K(cid:19)2 Z ∞ −∞(cid:18) ∂ ps dξ = 4c2(2KΓ + µ2)− 1 2 3(c − Fx) (cid:18)1 + π2 30(cid:19), (D.10) Z ∞ −∞ ∂ ps ∂K ∂ ps ∂c dξ = 2c2(2KΓ + µ2)1/2 3Γ(c − Fx)2 ×(cid:18)1 − 3Fx 2c − π2 15(cid:19), (D.11) ∂c (cid:19)2 Z ∞ −∞(cid:18) ∂ ps dξ = 2c2(2KΓ + µ2) 3Γ2(c − Fx)3 3 2 ×(cid:18) π2 ∂2 ps ∂ξ2 dξ = − Fx(c − Fx) − c2 15 c2(2KΓ + µ2)3/2 (cid:19), (D.12) , (D.13) 3Γ(c − Fx)3 Z ∞ −∞ ∂ ps ∂K ×(cid:18)1 + 2π2 105(cid:19), (D.17) Z ∞ −∞ ∂ ps ∂K ∂2 ps ∂K 2 dξ = − 2c2Γ 3(c − Fx)(2KΓ + µ2)3/2 Z ∞ −∞ ∂ ps ∂c ∂2 ps ∂ξ2 dξ = c2(2KΓ + µ2)5/2 15Γ2(c − Fx)4 ×(cid:18)1 + 2 c Fx(cid:19), c2(2KΓ + µ2)1/2 Γ(c − Fx) (D.14) , (D.15) psdξ = c2(2KΓ + µ2)3/2 3Γ2(c − Fx)2 (cid:18)1 − 2Fx c (cid:19),(D.16) dξ = 2c3(2KΓ + µ2)1/2 3Γ(c − Fx)2 ∂ ps ∂K −∞ Z ∞ Z ∞ Z ∞ ∂ ps ∂c −∞ −∞ psdξ = ∂K(cid:19)2 ps(cid:18) ∂ ps Z ∞ −∞ ps ∂ ps ∂c ∂ ps ∂K dξ = 2c3(2KΓ + µ2)3/2 9Γ2(c − Fx)3 ×(cid:18)1 − 2π2 35 − 2Fx c (cid:19), (D.18) c2(2KΓ + µ2)1/2 3Γ(c − Fx)2 dξ = Z ∞ −∞ ∂ ps ∂K ∂ ps ∂Fx Z ∞ −∞ ∂ ps ∂c ∂ ps ∂Fx ×(cid:18)1 + 2π2 15 (cid:19), dξ = c2(2KΓ + µ2)3/2 3Γ2(c − Fx)3 Fx 2π2 15 − ×(cid:18)1 − c (cid:19), Z ∞ −∞ ps ∂ ps ∂K ∂ ps ∂Fx dξ = 2c3(2KΓ + µ2)3/2 9Γ2(c − Fx)3 Z ∞ −∞ ps ∂ ps ∂c ∂ ps ∂Fx dξ = ×(cid:18)1 + 2π2 35 (cid:19), 2c3(2KΓ + µ2)5/2 15Γ3(c − Fx)4 2π2 4Fx 3c − ×(cid:18)1 − 21 (cid:19), (D.22) (D.19) (D.20) (D.21) Z ∞ −∞ Z ∞ −∞ ∂ ps ∂Fx(cid:19)2 ∂c (cid:18) ∂ ps ×(cid:20) 2π2 7 ∂ ps ∂K ∂ ps ∂c ∂ ps ∂Fx dξ = c3(2KΓ + µ2)5/2 45Γ3(c − Fx)5 2Fx c (cid:18)1 + 2π2 7 (cid:19)(cid:21), c3(2KΓ + µ2)3/2 − 1 − dξ = 9Γ2(c − Fx)4 2π2 ×(cid:20)1 − 2π2 35 − 2Fx c (cid:18)1 + 35 (cid:19)(cid:21), 15 (D.25) (D.26) ×(cid:18)1 + π2 30(cid:19), (D.27) c2(2KΓ + µ2)−1/2 3(c − Fx)2 2π2 Fx c (cid:18)1 + 15 (cid:19)(cid:21), + (D.28) Z ∞ −∞ ∂ ps ∂c ∂2 ps ∂K 2 dξ = π2 ×(cid:20)− 5 Z ∞ −∞ ∂ ps ∂K ∂2 ps ∂K∂Fx dξ = 2c2(2KΓ + µ2)−1/2 3(c − Fx)2 ×(cid:18)1 + π2 30(cid:19), (D.29) Z ∞ −∞ ∂ ps ∂c ∂2 ps ∂K∂Fx dξ = 2c2(2KΓ + µ2)1/2 ×(cid:20)1 − π2 6 − 3Γ(c − Fx)3 π2 Fx c (cid:18)1 − 15(cid:19)(cid:21), (D.30) Z ∞ −∞ ∂ ps ∂K ∂2 ps ∂F 2 x dξ = 2c2(2KΓ + µ2)1/2 3Γ(c − Fx)3 (cid:18)1 + π2 30(cid:19),(D.31) Z ∞ −∞(cid:18) ∂ ps ∂K(cid:19)2 ∂ ps ∂Fx Z ∞ −∞ ∂ ps ∂Fx(cid:19)2 ∂K(cid:18) ∂ ps dξ = c3(2KΓ + µ2)1/2 3Γ(c − Fx)3 ×(cid:18)1 + 2π2 21 (cid:19), dξ = c3(2KΓ + µ2)3/2 9Γ2(c − Fx)4 ×(cid:18)1 + 6π2 35 (cid:19), (D.23) Z ∞ −∞ ∂ ps ∂c ∂2 ps ∂F 2 x dξ = ×(cid:20)1 − π2 6 2c2(2KΓ + µ2)3/2 3Γ2(c − Fx)4 π2 Fx c (cid:18)1 − 15(cid:19)(cid:21). − (D.32) (D.24) Using these integrals, we obtain the CCEs (C.10) and (C.11). 16 [1] J. Folkman, Tumor angiogenesis. Adv. Cancer Res. 19, 331-358 (1974). [2] P. F. Carmeliet, Angiogenesis in life, disease and medicine. Nature 438, 932-936 (2005). [3] P. Carmeliet, and R.K. Jain, Molecular mechanisms and clinical applications of angiogenesis. Nature 473, 298-307 (2011). [4] W.D. Figg, and J. Folkman (eds), Angiogenesis. An In- tegrative Approach From Science to Medicine (Springer, Berlin 2008). [5] P. Carmeliet, and M. Tessier-Lavigne, Common mech- anisms of nerve and blood vessel wiring. Nature 436, 193-200 (2005). [6] R.F. Gariano, and T.W. Gardner, Retinal angiogenesis in development and disease. Nature 438, 960-966 (2005). [7] M. Fruttiger, Development of the retinal vasculature. An- giogenesis 10, 77-88 (2007). [8] L.A. Liotta, G.M. Saidel, and J. Kleinerman, Diffusion model of tumor vascularization. Bull. Math. Biol. 39, 117-128 (1977). [9] C. L. Stokes, and D. A. Lauffenburger, Analysis of the roles of microvessel endothelial cell random motility and chemotaxis in angiogenesis. J. Theoret. Biol. 152, 377- 403 (1991). [10] M.A.J. Chaplain, and A. Stuart, A model mechanism for the chemotactic response of endothelial cells to tumour angiogenesis factor. IMA J. Math. Appl. Med. Biol. 10, 149-168 (1993). [11] M.A.J. Chaplain, The mathematical modelling of tumour angiogenesis and invasion. Acta Biotheor. 43, 387-402 (1995). [12] A. R. A. Anderson, and M. A. J. Chaplain, Continuous and discrete mathematical models of tumor-induced an- giogenesis. Bull. Math. Biol. 60, 857-900 (1998). [13] S. Tong, and F. Yuan, Numerical simulations of angio- genesis in the cornea. Microvascular Research 61, 14-27 (2001). [14] H.A. Levine, S. Pamuk, B.D. Sleeman, and M. Nilsen- Hamilton, Mathematical modeling of the capillary forma- tion and development in tumor angiogenesis: penetration into the stroma. Bull. Math. Biol. 63, 801-863 (2001). [15] M. J. Plank, and B. D. Sleeman, Lattice and non-lattice models of tumour angiogenesis. Bull. Math. Biol. 66, 1785-1819 (2004). [16] N.V. Mantzaris, S. Webb, H.G. Othmer, Mathematical modeling of tumor-induced angiogenesis. J. Math. Biol. 49, 111-187 (2004). [17] S. Sun, M. F. Wheeler, M. Obeyesekere, and C. W. Patrick Jr., A deterministic model of growth factor- induced angiogenesis. Bull. Math. Biol. 67, 313-337 (2005). [18] S. Sun, M. F. Wheeler, M. Obeyesekere, and C. W. Patrick Jr., Multiscale angiogenesis modeling using mixed finite element methods. Multiscale Model Simul. 4, 1137 (2005). [19] A. St´ephanou, S. R. McDougall, A. R. A. Anderson, and M. A. J. Chaplain, Mathematical modelling of the in- fluence of blood rheological properties upon adaptative tumour-induced angiogenesis. Mathematical and Com- puter Modelling 44, 96-123 (2006). model exhibiting branching and anastomosis during tumor-induced angiogenesis. Biophys. J. 92, 3105-3121 (2007). [21] V. Capasso, and D. Morale, Stochastic modelling of tumour-induced angiogenesis. J. Math. Biol. 58, 219-233 (2009). [22] T. Jackson, and X. Zheng, A cell-based model of endothe- lial cell migration, proliferation and maturation during corneal angiogenesis. Bulletin of Mathematical Biology 72(4), 830-868 (2010). [23] A. Das, D.A. Lauffenburger, H. Asada, and R. D. Kamm, A hybrid continuum-discrete modelling approach to pre- dict and control angiogenesis: analysis of combinatorial growth factor and matrix effects on vessel-sprouting mor- phology. Phil. Trans. Roy. Soc. A 368, 2937-2960 (2010). [24] K.R. Swanson, R.C. Rockne, J. Claridge, M.A. Chaplain, E.C. Alvord Jr, and A.R.A. Anderson, Quantifying the role of angiogenesis in malignant progression of gliomas: in silico modeling integrates imaging and histology. Can- cer Res. 71, 7366-7375 (2011). [25] M. Scianna, L. Munaron, and L. Preziosi, A multiscale hybrid approach for vasculogenesis and related poten- tial blocking therapies. Prog. Biophys. Mol. Biol. 106(2), 450-462 (2011). [26] M. Scianna, J. Bell, and L. Preziosi, A review of math- ematical models for the formation of vascular networks. J. Theor. Biology 333, 174-209 (2013). [27] S.L. Cotter, V. Klika, L. Kimpton, S. Collins, and A. E. P. Heazell, A stochastic model for early placental devel- opment. J.R. Soc. Interface 11, 20140149 (2014). [28] E. Dejana and M.G. Lampugnani, Differential adhesion drives angiogenesis. Nature Cell Biol. 16, 305-306 (2014). [29] K. Bentley, C.A. Franco, A. Philippides, R. Blanco, M. Dierkes, V. Gebala, F. Stanchi, M. Jones, I.M. Aspalter, G. Cagna, S. Westrom, L. Claesson-Welsh, D. Vestweber, and H. Gerhardt, The role of differential VE-cadherin dynamics in cell rearrangement during angiogenesis. Nat. Cell Biol. 16(4), 309-321 (2014). [30] L.L. Bonilla, V. Capasso, M. Alvaro, and M. Carretero, Hybrid modeling of tumor-induced angiogenesis. Phys. Rev. E 90, 062716 (2014). [31] T. Heck, M.M. Vaeyens, H. Van Oosterwyck, Computa- tional models of sprouting angiogenesis and cell migra- tion: towards multiscale mechanochemical models of an- giogenesis. Math. Model. Nat. Phen. 10, 108-141 (2015). [32] F. Terragni, M. Carretero, V. Capasso, and L.L. Bonilla, Stochastic model of tumour-induced angiogenesis: En- semble averages and deterministic equations. Phys. Rev. E 93, 022413 (2016). [33] L.L. Bonilla, M. Carretero, F. Terragni, and B. Birnir, Soliton driven angiogenesis. Sci. Rep. 6, 31296 (2016). doi:10.1038/srep31296. [34] N. Manton and P. Sutcliffe, Topological solitons (Cam- bridge U.P., Cambridge UK 2004). [35] M. Remoissenet, Waves called solitons: Concepts and ex- periments, 3rd ed. (Springer, Berlin 1999). [36] A. Carpio and G. Duro, Well posedness of an integrodif- ferential kinetic model of Fokker-Planck type for angio- genesis. Nonlinear Analysis: Real World Applications 30, 184-212 (2016). [20] A.L. Bauer, T.L. Jackson, and Y. Jiang, A cell-based [37] L. L. Bonilla, and S.W. Teitsworth, Nonlinear Wave Methods for Charge Transport (Wiley-VCH, Weinheim, 2010). [38] F. G. Mertens, H. J. Schnitzer, and A. R. Bishop, Hier- archy of equations of motion for nonlinear coherent ex- citations applied to magnetic vortices. Phys. Rev. B 56, 2510-2520 (1997). [39] B. S´anchez-Rey, N. R. Quintero, J. Cuevas-Maraver, and M. A. Alejo, Collective coordinates theory for discrete soliton ratchets in the sine-Gordon model. Phys. Rev. E 90, 042922 (2014). [40] C. W. Gardiner, Stochastic methods. A handbook for the natural and social sciences, 4th ed (Springer, Berlin 2010). [41] A. Carpio, G. Duro, and M. Negreanu, Constructing so- lutions for a kinetic model of angiogenesis in annular do- mains. Preprint arXiv:1612.07389, to appear in Applied Mathematical Modelling. doi:10.1016/j.apm.2016.12.028 17
1703.02536
1
1703
2017-03-07T14:23:39
Generic transport mechanisms for molecular traffic in cellular protrusions
[ "physics.bio-ph" ]
Transport of molecular motors along protein filaments in a half-closed geometry is a common feature of biologically relevant processes in cellular protrusions. Using a lattice gas model we study how the interplay between active and diffusive transport and mass conservation leads to localised domain walls and tip localisation of the motors. We identify a mechanism for task sharing between the active motors (maintaining a gradient) and the diffusive motion (transport to the tip), which ensures that energy consumption is low and motor exchange mostly happens at the tip. These features are attributed to strong nearest-neighbour correlations that lead to a strong reduction of active currents, which we calculate analytically using an exact moment-identity, and might prove useful for the understanding of correlations and active transport also in more elaborate systems.
physics.bio-ph
physics
Generic transport mechanisms for molecular traffic in cellular protrusions Isabella R. Graf1 and Erwin Frey1, ∗ 1Arnold-Sommerfeld-Center for Theoretical Physics and Center for NanoScience, Department of Physics, Ludwig-Maximilians-Universitat Munchen, D -- 80333 Munich, Germany Transport of molecular motors along protein filaments in a half-closed geometry is a common feature of biologically relevant processes in cellular protrusions. Using a lattice gas model we study how the interplay between active and diffusive transport and mass conservation leads to localised domain walls and tip localisation of the motors. We identify a mechanism for task sharing between the active motors (maintaining a gradient) and the diffusive motion (transport to the tip), which ensures that energy consumption is low and motor exchange mostly happens at the tip. These features are attributed to strong nearest-neighbour correlations that lead to a strong reduction of active currents, which we calculate analytically using an exact moment-identity, and might prove useful for the understanding of correlations and active transport also in more elaborate systems. Linear protrusions of cells, as for instance filopodia or stereocilia, perform multiple tasks in living organisms, ranging from cell migration and signal transduction to wound healing. They contain actin filaments cross-linked into bundles by actin-binding proteins [1 -- 4], and molec- ular motors of the myosin family which interact with actin filaments and walk on them in a persistent, uni- directional fashion towards the tip of the protrusion [3 -- 9]. These motors play an important role in the biological function of protrusions [1, 4, 5]. In particular, they are known to localise to the tips of filopodia and stereocilia, and are (jointly) responsible for length control [2 -- 12]. Motivated by these observations, various models have been investigated. Some are detailed mathematical mod- els addressing specific biological issues. These include the role of motor transport in shaping the concentration profile of G-actin at the base of protrusions [13], the lo- calisation of different proteins along stereocilia [14], and the effect of myosin X on filopodial growth [15]. Comple- mentary, simplified conceptual models have been stud- ied asking how the interplay between active and diffusive transport in open systems may lead to non-equilibrium phase transitions and ensuing steady states with interest- ing correlations and nontrivial density profiles [16 -- 20]. The latter are based on the totally asymmetric simple exclusion process (TASEP) [21, 22], a lattice-gas model that, despite its simple structure, has become a paradigm for non-equilibrium dynamics [23, 24]. Here we present and analyse a conceptual model cap- turing two basic properties of the motion of persistent, plus-end directed motors inside narrow, elongated cellu- lar protrusions. First, there is an interplay between two genuinely different types of dynamics: directed (active) transport with steric hindrance along actin filaments, and diffusive motion in the cytoplasm. These are coupled by particle exchange between the filament and the cyto- plasm. Second, the half-closed geometry of cellular pro- trusions is special: At one end, the protrusions are con- nected to the cell body and thus to a reservoir, whereas everywhere else protein diffusion is confined by the cell membrane, so that mass conservation and resource lim- FIG. 1. Illustration of the two-lane lattice-gas model com- prised of a TASEP and a SSEP with hopping rates ν and , respectively. The lanes are coupled via a symmetric attach- ment and detachment rate ω (cid:28) ν. In model A, detachment from the last site is given by β (cid:29) ω, while in model B exchange between the lanes is fully symmetric. At the tip of the sys- tems mass conservation holds, and at the base particles can enter the SSEP lane at rate α and exit at rate . itation play an important role there. The combination of mass-conservation (closure) on the one hand and the interplay of equilibrium (diffusion) and non-equilibrium (active transport) processes on the other hand is intrin- sically interesting to study as closure in a system entails no-flux boundary conditions that oppose currents from active transport. Here, we want to examine the inter- play of these mechanisms with the help of an abstract model that is motivated biologically but has a level of description that makes it possible to understand all the processes accounted for. We identify generic mechanisms based on correlations and non-equilibrium physics that could be of importance for biological systems as cellu- lar protrusions but, inevitably, predictions for biological systems are qualitative. Specifically, we consider a two-lane lattice-gas model in a half-closed geometry [Fig. 1] in steady state, simi- lar to Ref. [25]. One lane represents the actin filament and the second lane the cytoplasm. While there is a rich literature on the non-equilibrium dynamics of two- lane systems [20, 25 -- 43], very few of these studies ad- dress how the physics of non-equilibrium steady-states is affected by a half-closed geometry [25, 42]. We are interested in the limit where the actin filament and the cytoplasm are coupled weakly by attachment and detach- ment processes, while Ref. [25] focuses on the strong cou- ανεεεωωωβ01L2ωA:LωωB: pling limit. Our analyses show that, due to the closure of the system at the tip, there is only one type of density profile, namely, domain walls (DW) separating a high- and a low-density region. The limit where the width of the high-density region becomes microscopically small (of the order of a few lattice sites) corresponds to tip local- isation. Furthermore, correlations in such systems have not received much attention, and we want to investigate nearest-neighbour (NN) correlations on the filament. In a biological context, this is related to efficient transport on actin filaments and to the significance of steric hindrance of motors there. We find that those correlations reach high values close to the DW, and the transport current along the filament is strongly reduced compared to its mean-field (MF) prediction. This suggests an important role for the cytoplasm, namely to transport the proteins to the tip. Conversely, active transport effectively sets up and maintains a gradient of motor proteins. Our model consists of two coupled sublattices [Fig. 1], namely a TASEP and a SSEP (symmetric simple exclu- sion process) lane of L + 1 sites ∈{0, 1, ..., L}. The dy- namics on lane 1 (TASEP) are governed by a rate ν at which particles jump forward one site towards the right (tip) provided that the site in front of them is empty (exclusion). This corresponds to the directional motion of the motors on the filament that is oriented towards the tip. By convention, we measure time in units of ν (i.e. set ν = 1) and length in units of the system size (i.e. the lattice spacing is a = 1 L and the total length La = 1). In lane 2 (SSEP) particles jump non-directionally be- tween neighbouring sites, at rate  again respecting ex- clusion. This represents the diffusive motion in the cy- toplasm that is taken to be effectively one-dimensional in the thin cylinder-like protrusions (the steady-state be- haviour of a system with several lanes for diffusion ar- ranged on a cylinder around the TASEP can be reduced to the steady-state behaviour of this model [44]). Par- ticles enter or exit the system only at site 0 of lane 2 (base) but not at site 0 of lane 1. At rate α a parti- cle is injected provided the site is empty and at rate  a particle leaves the system. This reflects the exchange of motors between the protrusion base and the cell body. In the bulk both lanes are coupled via a rate ω at which particles jump from site i∈{0, 1, ..., L − 1} of lane 1 to site i of lane 2 or vice versa (each respecting exclusion). Since the biochemistry at the tip is only poorly under- stood, at site L we consider two extreme cases: In model A, particles jump from site L of lane 1 to site L of lane 2 at rate β (cid:29) ω (respecting exclusion) but not in the op- posite direction. This describes a scenario where motors at the tip detach mainly due to the lack of a filament subunit in front of them. In model B, the exchange rates between the lanes at sites L are the same as in the bulk. The comparison of both models stresses that seemingly minor changes in a non-equilibrium system may have a strong influence on the dynamics [45], and highlights the 2 FIG. 2. Representative steady-state DW profile ρT(x) (left) with covariances Cov(x) (right) on the TASEP lane are shown exemplarily for model A with L = 50, β = 0.2, Ω = 0.001, α = 0.1 and  = 0.025. The covariances are non-zero only in the vicinity of the DW. Using the fluctuation-corrected pro- file from DW theory, our predictions (dotted green curves) fit the simulation result (filled red circles) very well. If we use the refined MF profile (dot-dashed blue) instead of a step function (black dashed curve), the width of the DW and the strength of the covariances are strongly underestimated [44]. relevance of the biochemistry at the filament tip for the motor density profile. In the following, when considering the continuum limit a → 0, we focus on the mesoscopic limit for ω [18, 19], i.e. we keep Ω = ω/a fixed, thus en- suring the number of jumps between lanes is of the same order as that of entry or exit events (persistent motors) and implementing weak coupling between the diffusive and the directed motion. For simplicity, we take the at- tachment and detachment rates to be equal. However, the qualitative results do not change for different attach- ment and detachment rates ωA (cid:54)= ωD as long as both are still taken in the mesoscopic limit (not shown here). A single TASEP exhibits three phases, namely a max- imal current (MC), a low- (LD) and a high-density (HD) phase [16]. Moreover, on the phase boundary between the LD and HD phase the steady-state profile is given by a DW that performs a random walk. Due to the closure at the tip, we do not find a MC phase in our system [44], but instead observe localised DWs [17, 18]. That is, the generic steady-state TASEP profile ρT(x) for both mod- els is given by a localised DW separating low density at the base from high density at the tip [Fig. 2]. For generic parameters, the filament current is thus comparatively small, and restricted to a small part of the system. This might be beneficial from a biological point of view, since every motor step on the filament consumes ATP. The po- sition of the DW depends on the model parameters, and is shifted towards the tip (base) for smaller (higher) val- ues of σ := α α+ [44]. This ratio can be thought of as the motor density in the cell body, and thus the value for the cytoplasmic density at the protrusion base. The width of the DW decreases with increasing L and with increasing distance from the tip [44]. Generically, when describing the DW as a random walker (see below), we observe that its motion is mainly confined to a small part of the sys- tem. Hence we assume for the moment that the DW is fully localised and adopt a step-function ansatz for ρT(x) that holds to lowest order in a: ρT(x) = 0 for x∈ [0, z[, and ρT(x) = 1 for x∈ ]z, 1]. First, we determine the po- i (cid:105) = Prob{nµ i ∈{0, 1}, we have (cid:104)nµ sition z of the DW to find out whether tip localisation (i.e. z ≈ 1) occurs for certain parameter ranges or not. For this purpose we derive a mass-balance equation re- lating the total average occupancies of the TASEP (T) and the SSEP (S). We denote by nµ i the occupation num- ber on site i∈{0, . . . , L} of lane µ∈{T, S}, i.e. we write i = 0 if site i of lane µ is empty and nµ nµ i = 1 if it is oc- i = 1}; cupied. Since nµ ensemble averages are denoted by (cid:104) (cid:105). The steady-state condition in the bulk corresponds to a flux balance [44]: i−1 − fi−1)− (ρT (ρT i ) with the correla- i (cid:105). i =(cid:104)nµ tor fi =(cid:104)nT The difference of the TASEP currents, from site i− 1 to site i and from site i to site i + 1, (left-hand side) must equal the current between sites i of the TASEP and SSEP, ω((cid:104)nT i )(cid:105)), (right-hand side). i )(cid:105)−(cid:104)nS It follows that fi = f0 + ρT j ). At the base, there is no direct flux from the cell body into the 0 − filament; so the boundary condition is f0 = ρT ρS 0) and we find the following exact moment-identity i+1(cid:105) and the average occupancy ρµ 0 + ω(cid:80)i i (1− nT i − ρT i − fi) = ω(ρT i (1− nS 0 + ω(ρT i − ρS j − ρS j=1(ρT i nT fi = ρT i + ω j − ρS (ρT j ) . (1) i(cid:88) j=0 The systems are both closed at the tip but, due to the different attachment and detachment behaviour, the L)(cid:105), boundary conditions read fL−1 = ρT and fL−1 = ρT L) for models A and B, re- spectively. Combining these with Eq. (1), the following exact mass-balance equations can be derived for model A L−1 − β(cid:104)nT L(1−nS L−1 + ω(ρS L − ρT L −(cid:10)nT LnS L (cid:11)(cid:1) , (2) L−1(cid:88) (cid:0)ρT (cid:1) = −β(cid:0)ρT and similarly for model B: ω(cid:80)L j − ρS j=0 ω j j −ρS j=0(ρT j ) = 0. These equations relate the average occupancy on the two lanes in such a way that the total influx into the TASEP lane equals the total outflux from it [46]. Interestingly, these equations reveal that a global detailed balance holds for the total exchange between the two lanes, rather than local detailed balance for any pair of sites (cf. adsorption equilibrium in Ref. [25]). The moment-identity and the mass-balance equations are useful in two ways: (i) By using the DW ansatz, we are able to find an analytic formula for the DW position. (ii) The moment-identity enables us to express covariances with respect to NNs on the TASEP lane in terms of densities. To address the first issue, we determine the aver- age density ρS(x) on the SSEP lane corresponding to the fully-localised-DW ansatz by solving the bulk equa- tion for the SSEP. For that, we implement the contin- uum limit and, for model A, assume (cid:104)nT L [44]. Note, that we only need this MF approximation for the tip densities of model A. This is due to the oth- erwise symmetric attachment and detachment rates L(cid:105)≈ ρT LnS LρS 3 and the diffusive motion in the cytoplasm, for both of which the correlations drop out in the dynami- cal equation. With the resulting equation for ρS(x) we can then conclude z = 1− l cosh −1 [σ cosh(1/l)] and z = 1− l sinh −1 [σ sinh(1/l)] for models A and B, respec- tively, where l :=(cid:112)D/ω with D := a2 being the diffu- sion constant in the cytoplasm [44]; l can be understood as the typical dwell length for motors in the cytoplasm before attaching to the filament. Comparing these ex- pressions with our stochastic simulation results [47] we find excellent agreement [Fig. 3]. The phase diagrams for the two models are qualita- tively different. For model A, one can switch between the DW phase (z (cid:28) 1) and the tip-localisation phase (z ≈ 1) In con- by only slightly increasing the dwell length l. trast, for model B, tip localisation is attained only as σ → 0, even for large l [44]. This can be understood from the symmetry between the attachment and detachment processes in model B, which should also be reflected in a symmetry in occupancy between the filament and the cy- toplasm. For large l, the cytoplasm becomes well-mixed with constant density σ. So, the average density on the filament is σ as well, which is realised for a DW position z = 1− σ. For model A, fast diffusion in the cytoplasm leads to rapid diffusion of the motors away from the tip. Therefore, the exit rate at the tip is high and motors quickly leave the tip of the filament, thus enabling tip localisation but reducing large jamming. For small but finite σ, there is tip localisation also if the typical dwell length is smaller than the system size, l ≤ 1. That is, even if the motors have a relatively small D, tip locali- sation can occur if the tip has an enhanced detachment rate (model A). For model B, even by increasing l well beyond the system size, tip localisation occurs only for very low motor density σ at the base. These features may have interesting biological implica- tions: A higher detachment rate at the actin filament's end would promote tip localisation, but simultaneously avoid jamming. As a result, motor exchange between fila- ment and cytoplasm occurs primarily near the tip (where the DW is located) and the motors at the tip can be con- tinuously replenished by new ones delivered through the cytoplasm. Furthermore, energy consumption (ATP hy- drolysis) is low, as the filament current is kept small and mainly restricted to the tip area. Transport to the tip is facilitated mainly by diffusion in the cytoplasm, which does not consume chemical energy. In summary, energy from ATP hydrolysis could efficiently be used to localise motors to the filament tip, while material transport is facilitated by diffusion in the cytoplasm [44]. In this regard, the model is also interesting from a theoretical point of view, as it allows for the calculation i − fi that depends on NN of the filament current J T correlations or, equivalently, the NN covariances i = ρT Covi := (cid:104)nT i+1(cid:105) − ρT i+1 = fi − ρT i ρT i nT i ρT i+1 (3) 4 i (1− ρT i+1)− J T i = ω(cid:80)i The j=0(ρS covariances j − ρT i+1=1nT are i =1nT i =1}≥ Prob{nT non-negative i+1=1}≥ Prob{nT 0.2. We have Covi = ρT i , where the first term corresponding to the MF current dominates. There- fore, the actual current J T j ) can be orders of magnitude smaller than the MF current, and scales with the particle exchange rate ω [44]. That is, density correlations dominate in such a way that they suppress the TASEP current substantially. This demon- strates that, even if basic properties of the single TASEP are captured well by MF theory, correlation effects in TASEP-based systems should be studied more closely and might lead to unanticipated features [50 -- 52]. everywhere, Covi ≥ 0, so the conditional probabilities obey the i =1} inequalities Prob{nT and i+1=1} at any site, Prob{nT im- plying that particles typically form clusters. As a result, from the DW region onwards (where the covariances are highest), the mean time a particle spends at a certain site is increased considerably compared to freely moving motors. This is due to the effective jump rate on the filament, which is decreased by excluded-volume effects. Thus, in case of tip localisation, particles spend more time near the tip than in the main part of the filament. This prolongation of the residence time is further enhanced by the exclusion in the cytoplasm that prevents motors from detaching if the cytoplasmic tip density is high. This is especially important for model A, where the cytoplasmic tip density takes a value of 1, which is much higher than in the bulk [44]. Biologically, the extended residence time at the tip might facilitate the tasks of the motors or their cargo at the tip. Our results all essentially rely on the exact moment- identity and the exact mass-balance equations. The derivation of both depends on the TASEP dynamics and the coupling between the two lanes, but not at all on the dynamics of the second (here SSEP) lane. Hence, those equations do not change if this dynamics on the second lane is modified, and we expect them to be useful for other model systems in which a TASEP lane is coupled to another lattice via attachment and detachment kinet- ics. To our best knowledge, the moment-identity has not been mentioned before. We believe that it might open doors in the understanding of correlations and the pre- diction of active currents also in more elaborate models. Furthermore, both equations could easily be generalised to the case of different attachment and detachment rates. Our models could also be varied in other interesting ways. One could account for the three-dimensional ge- ometry or for polymerisation and depolymerisation of the filament and the accompanying changes in length. Nev- ertheless, we expect that some of the phenomena seen here should be robust against such modifications. Tip localisation, which is mostly based on mass conservation at the tip, should still be present. Second, the TASEP current might still be suppressed, and the roles of the TASEP and the diffusive lane in being responsible for FIG. 3. Phase diagrams for the DW position z (colour-coded) for L = 50 and Ω = 0.001 for model A (upper panel, β = 0.2) and model B (lower panel) as a function of the typical dwell length l in the cytoplasm and the cytoplasmic density at the protrusion base σ: the simulation results are shown on the left, the theoretical predictions for the DW position are shown on the right. The black thick lines in the diagrams for model A show the phase boundary z = 1 as obtained from theory. The dashed black lines in the diagrams obtained from theory are, from right to left respectively, contour lines of constant z = 0.2, 0.4, 0.6 and additionally z = 0.8 for model B. To go beyond MF, we use on the TASEP lane. Eq. (1), which relates fi to the average densities, and find Covi = ρT j ), which in the continuum limit translates to i+1) + ω(cid:80)i (cid:90) x Cov(x) = ρT(x)(cid:2)1−ρT(x+a)(cid:3) + Ω dy (ρT−ρS) . (4) i (1− ρT j − ρS j=0(ρT 0 The value of the first term in Eq. (4) depends sensitively on the width and shape of the DW (the density profile ρT(x) is increasing with x, so that ρT(x)(1− ρT(x+a)) is maximal if both ρT(x) and ρT(x+a) are close to 0.5). Therefore, one needs to refine the fully-localised-DW ansatz, by taking into account the stochastic dynamics of the DW position. Following Refs. [48, 49] we consider the DW as a random walker with (site-dependent) hopping rates depending on currents and densities in the low- and high-density regions. For small a we find , (5) ρT(x) ≈ erf ((x − z)/W (z)) + erf (z/W (z)) erf ((1 − z)/W (z)) + erf (z/W (z)) where W (z) =(cid:112)2σal sinh (z/l) [44]. This fluctuation- corrected DW profile, as well as the covariance obtained from it, agree very well with our simulation data [Fig. 2]. If one uses a refined MF method instead, account- ing for second-order spatial derivatives, the DW width and the strength of the correlations are both markedly underestimated [44]. In general, covariances are non-zero only close to the DW, but there they can reach quite high values of around simulationtheorymodel Amodel B tip localisation and motor transport, respectively, should remain untouched. This seems to be supported by more elaborate models in a related context [13, 15]. We thank Matthias Rank, Louis Reese, and Emanuel Reithmann for critical reading of this manuscript and for helpful discussions. This research was supported by the German Excellence Initiative via the program "NanoSys- tems Initiative Munich" (NIM) and by the Deutsche Forschungsgemeinschaft (DFG) through the Graduate School of Quantitative Biosciences Munich (QBM). 5 (2007). [24] T. Chou, K. Mallick, and R. K. P. Zia, Rep. Prog. Phys. 74, 116601 (2011). [25] M. Muller, S. Klumpp, and R. Lipowsky, J. Phys.: Con- dens. Matter 17, S3839 (2005). [26] S. Klumpp and R. Lipowsky, J. Stat. Phys. 113, 233 (2003). [27] V. Popkov and G. M. Schutz, J. Stat. Phys. 112, 523 (2003). [28] E. Pronina and A. B. Kolomeisky, J. Phys. A 37, 9907 (2004). [29] B. Schmittmann, J. Krometis, and R. K. P. Zia, Euro- phys. Lett. 70, 299 (2005). [30] E. Pronina and A. B. Kolomeisky, Physica A 372, 12 (2006). ∗ Correspondence please to [email protected]. [1] C. Revenu, R. Athman, S. Robine, and D. Louvard, Nat. [31] T. Reichenbach, E. Frey, and T. Franosch, New J. Phys. 9, 159 (2007). [32] E. Pronina and A. B. Kolomeisky, J. Phys. A 40, 2275 Rev. Mol. Cell Biol. 5, 635 (2004). (2007). [2] P. K. Mattila and P. Lappalainen, Nat. Rev. Mol. Cell [33] R. Jiang, R. Wang, and Q.-S. Wu, Physica A 375, 247 Biol. 9, 446 (2008). (2007). [3] F. Les Erickson, A. C. Corsa, A. C. Dos´e, and B. Burn- [34] T. Reichenbach, T. Franosch, and E. Frey, Eur. Phys. J. side, Mol. Biol. Cell 14, 4173 (2003). E 27, 47 (2008). [4] R. Nambiar, R. E. McConnell, and M. J. Tyska, Cell. [35] R. Wang, M. Liu, and R. Jiang, Physica A 387, 457 Mol. Life Sci. 67, 1239 (2010). (2008). [5] F. T. Salles, R. C. Merritt, U. Manor, G. W. Dougherty, A. D. Sousa, J. E. Moore, C. M. Yengo, A. C. Dose, and B. Kachar, Nat. Cell Biol. 11, 443 (2009). [6] M. A. Hartman and J. A. Spudich, J. Cell Sci. 125, 1627 (2012). [36] M. R. Evans, P. A. Ferrari, and K. Mallick, J. Stat. Phys. 135, 217 (2009). [37] C. Schiffmann, C. Appert-Rolland, and L. Santen, J. Stat. Mech. 2010, P06002 (2010). [38] A. Melbinger, T. Reichenbach, T. Franosch, and E. Frey, [7] M. L. Kerber and R. E. Cheney, J. Cell Sci. 124, 3733 Phys. Rev. E 83, 031923 (2011). (2011). [39] B. Saha and S. Mukherji, J. Stat. Mech. 09, P09004 [8] T. Kambara, S. Komaba, and M. Ikebe, J. Biol. Chem. (2013). 281, 37291 (2006). [40] A. K. Gupta and I. Dhiman, Phys. Rev. E 89, 022131 [9] J. E. Bird, Y. Takagi, N. Billington, M.-P. Strub, J. R. Sellers, and T. B. Friedman, Proc. Nat. Acad. Sci. USA 111, 12390 (2014). (2014). [41] D. Johann, D. Goswami, and K. Kruse, Phys. Rev. E 89, 042713 (2014). [10] A. K. Rzadzinska, M. E. Schneider, C. Davies, G. P. Ri- [42] I. Pinkoviezky and N. S. Gov, Phys. Rev. E 89, 052703 ordan, and B. Kachar, J. Cell Biol. 164, 887 (2004). (2014). [11] U. Manor, A. Disanza, M. Grati, L. Andrade, H. Lin, P. P. D. Fiore, G. Scita, and B. Kachar, Curr. Biol. 21, 167 (2011). [12] I. A. Belyantseva, E. T. Boger, and T. B. Friedman, [43] A. I. Curatolo, M. R. Evans, Y. Kafri, and J. Tailleur, J. Phys. A 49, 095601 (2016). [44] See Supplemental Material at [url will be inserted by pub- lisher]. Proc. Nat. Acad. Sci. USA 100, 13958 (2003). [45] L. Reese, A. Melbinger, and E. Frey, Interface Focus 4 [13] P. Zhuravlev, Y. Lan, M. Minakova, and G. Papoian, (2014), 10.1098/rsfs.2014.0031. Proc. Nat. Acad. Sci. USA 109, 10849 (2012). [14] M. Naoz, U. Manor, H. Sakaguchi, B. Kachar, and N. S. Gov, Biophys. J. 95, 5706 (2008). [15] K. Wolff, C. Barrett-Freeman, M. R. Evans, A. B. Gory- and D. Marenduzzo, Phys. Biol. 11, 016005 achev, (2014). [16] J. Krug, Phys. Rev. Lett. 67, 1882 (1991). [17] R. Lipowsky, S. Klumpp, and T. M. Nieuwenhuizen, Phys. Rev. Lett. 87, 108101 (2001). [46] S. Klumpp, M. J. I. Muller, and R. Lipowsky, "Traffic and Granular Flow '05" (Springer, Berlin, 2007). [47] D. T. Gillespie, J. Comput. Phys. 22, 403 (1976). [48] B. Derrida, M. R. Evans, and K. Mallick, J. Stat. Phys. 79, 833 (1995). [49] A. B. Kolomeisky, G. M. Schutz, E. B. Kolomeisky, and J. P. Straley, J. Phys. A 31, 6911 (1998). [50] V. Popkov, A. R´akos, R. D. Willmann, A. B. Kolomeisky, and G. M. Schutz, Phys. Rev. E 67, 066117 (2003). [18] A. Parmeggiani, T. Franosch, and E. Frey, Phys. Rev. [51] K. Tsekouras and A. B. Kolomeisky, J. Phys. A 41, Lett. 90, 086601 (2003). 095002 (2008). [19] A. Parmeggiani, T. Franosch, and E. Frey, Phys. Rev. E [52] E. Reithmann, L. Reese, and E. Frey, Phys. Rev. Lett. 70, 046101 (2004). 117, 078102 (2016). [20] M. Evans, Y. Kafri, K. Sugden, and J. Tailleur, J. Stat. Mech. 06, P06009 (2011). [21] C. T. MacDonald, J. H. Gibbs, and A. C. Pipkin, Biopolymers 6, 1 (1968). [22] F. Spitzer, Adv. Math. 5, 246 (1970). [23] R. A. Blythe and M. R. Evans, J. Phys. A 40, R333 1 Generic transport mechanisms for molecular traffic in cellular protrusions Supplemental Material: This Supplemental Material gives details on the mathematical analysis of the lattice gas model. We will explain more thoroughly why the generic steady-state TASEP density profile is given by a domain wall and why there is no maximal current phase. Furthermore, the calculation of the density profiles both for TASEP and SSEP are shown explicitly and the analytic expressions for the position of the domain wall are derived. It is demonstrated how domain wall theory is used concretely to improve on the mean-field TASEP density profiles, and on the prediction for the covariances. Those are relevant to see how the actual TASEP current differs from the expected mean-field current. We also display a comparison of the different currents (on the TASEP, on the SSEP and those for attachment and detachment) for a parameter set of model A where tip localisation occurs. Finally, we show that the steady-state behaviour of a geometry with several lanes for diffusion arranged on a cylinder around the TASEP lane can be reduced to the steady-state behaviour of our model by scaling the parameters for diffusion and attachment/detachment with the number of lanes for diffusion. The same holds true for a model where instead of exclusion on the lane for diffusion a finite carrying capacity Nmax > 1 is used. CALCULATION OF THE DENSITY PROFILES To set the stage, let us denote by nµ write nµ equations of the Markov processes, corresponding to model A and B as shown in Fig. 1, are given by i = 0 if site i of lane µ is empty and nµ i the occupation number on site i ∈ {0, . . . , L} of lane µ ∈ {T, S}, i.e. we will i = 1 if it is occupied (T: TASEP, S: SSEP). The common bulk master dProb{nT i } dProb{nS i } dt dt (cid:16) (cid:16) = Prob{nT i−1, nT =  + ω Prob{nS Prob{nT i } − Prob{nT i , nT i } + Prob{nS i } − Prob{nS i , nT i }(cid:17) i+1, nS , i+1} + ω Prob{nS i } − Prob{nS i−1, nS i , nS (cid:16) i , nT i } − Prob{nT i+1} − Prob{nS i , nS i , nS i , nS + i }(cid:17) i−1}(cid:17) , i , nT i = 1 and Prob{nµ i } denotes the probability that nµ i }− Prob{nT where Prob{nµ j = 0. The term i+1} is due to the jump process on the TASEP lane that respects the exclusion property Prob{nT i−1, nT and occurs at bare rate ν = 1. The terms proportional to ω describe the exchange between the lanes, again respecting the exclusion. And finally, the term proportional to  describes the diffusion on the SSEP lane. Note that we assume exclusion not only on the filament but also in the cytoplasm. This is based on the idea that, due to the finite size of particles, there should be a maximal number inside any finite volume element, introducing a carrying capacity Nmax. If we assume that the maximal effective attachment and detachment rate stay the same, i.e. if we assume that attachment happens at rate ω(cid:0)nS(cid:0)1 − nT(cid:1)(cid:1) and detachment at rate ω(cid:0)nT(cid:0)Nmax − nS(cid:1)(cid:1) where nS ∈ {0, 1, . . . , Nmax}, j} the one that nµ i = 1 and nν i , nν the case Nmax finite but arbitrary can be reduced to Nmax = 1 by redefinition of the parameters, and in the following we will focus only on the case Nmax = 1. We will, however, come back to this case again in the last paragraph of the Supplemental Material when we discuss the case of several lanes for diffusion. (cid:104)nµ occupied by one particle/motor: The bulk master equations can be rewritten in terms of averages over the occupation numbers by using that i (cid:105) = Prob (nµ i } hold, as each site can be either empty or i = 1) = Prob{nµ i )(cid:105) = Prob{nµ i } and (cid:104)nµ i (1 − nν i , nν where fi =(cid:10)nT i nT i+1 (cid:11) is the nearest-neighbour correlator for the TASEP. Summing both equations we find i−1 − 2ρS i i = ρT i = (cid:0)ρS i−1 − fi−1 − ρT i+1 + ρS ∂tρT ∂tρS ∂t(ρT i + ρS i ) = ρT i−1 − fi−1 − ρT i+1 + ρS for the time evolution of the combined density ρT i in the bulk. At the left boundary (base) we find (cid:1) , (cid:1) , i − ρT i − ρS i i i + fi + ω(cid:0)ρS (cid:1) + ω(cid:0)ρT i + fi + (cid:0)ρS 0 + (cid:0)ρS (cid:1) 1 − ρS 0 − ρT 0 0 (cid:1) i i−1 − 2ρS (cid:1) 0 − ρS 0 (cid:1) + ω(cid:0)ρT i + ρS 0 + f0 + ω(cid:0)ρS 0) − ρS 0 = −ρT 0 = α(1 − ρS ∂tρT ∂tρS (S1) (S2) (S3) (S4) for both models. For the TASEP lane there is only outflux from site 0 to site 1 or exchange with site 0 of the SSEP lane. On the SSEP lane, there is influx at rate α (respecting the exclusion at site 0 of the SSEP), outflux with the diffusion rate , diffusion between site 0 and site 1 and exchange with site 0 of the TASEP. The behaviour at the right boundary (tip) differs between the models and is given by for model A, and by ∂tρT ∂tρS (cid:11)(cid:1) (cid:11)(cid:1) L = ρT L−1 − ρS L−1 − fL−1 − β(cid:0)ρT L −(cid:10)nT (cid:1) + β(cid:0)ρT L −(cid:10)nT L−1 − fL−1 + ω(cid:0)ρS (cid:1) + ω(cid:0)ρT L = (cid:0)ρS L = (cid:0)ρS (cid:1) L − ρT L − ρS L−1 − ρS L = ρT ∂tρT ∂tρS L L L L LnS L LnS L (cid:1) 2 (S5) (S6) (S7) (S8) (S9) (S10) for model B. For model A, at site L of the TASEP lane there is influx from the neighbouring site L − 1 and outflux to site L of the SSEP lane at bare rate β respecting the exclusion. For site L of the SSEP lane there is diffusion between sites L − 1 and L of the SSEP lane and influx from site L of the TASEP lane. For model B, we have the same behaviour except that the asymmetric exchange between sites L of the TASEP and SSEP lane is replaced by (cid:1). Since the exchange terms drop out when considering the time derivative of the symmetric exchange ω(cid:0)ρT L − ρS L ρT L + ρS L, at the tip of both models it holds ∂t(ρT L + ρS L) = ρT L−1 − ρS L−1 + J S L−1, L−1 − fL−1 + (cid:0)ρS 0 + (cid:0)ρS 0) − ρS L (cid:1) = J T (cid:1) = −J T and at the base ∂t(ρT 0 + ρS 0) = −ρT 0 + f0 + α(1 − ρS 1 − ρS 0 0 + α(1 − ρS 0) − ρS 0 − J S 0 where we introduced the local currents and J T i = ρT i − fi i = (cid:0)ρS J S i − ρS i+1 (cid:1) for i = 0, . . . , L − 1 on the filament and in the cytoplasm, respectively. With that Eq. (S3) translates to ∂t(ρT i + ρS i ) = J T i−1 + J S (S11) L ≡ 0, Note that we can introduce currents J T/S−1 and J T/S and since there is no direct influx from the left into the TASEP J T−1 = 0. In order to be consistent with the structure of Eq. (S11) where the currents appear as Ji−1 − Ji, we define J S−1 = α(1 − ρS as well. However, as the system is closed at the tip J T/S 0) − ρS (S12) L 0 i−1 − J T i − J S i . so that ∂t(ρT 0 + ρS 0) = −J T 0 + J S−1 − J S 0 holds. As mentioned above, we are interested in the steady-state behaviour of the system and therefore set ∂tρT /S i ≡ 0 for all i. In particular, we have 0 = ∂t L(cid:88) (ρT i + ρS i ) = J S−1 where we used Eqs. (S7), (S8), (S11). As a result, i=0 ρS 0 = α α +  := σ, (S13) and the parameter σ as used for the phase diagrams in Fig. 3 can be identified as the motor density in the cell body. To proceed, let us now use the continuum limit where a = 1 L → 0. That is, we will replace the discrete lattice by the continuous space [0, 1] and the average occupation numbers are replaced by a continuous density i → ρµ (x = xi) where xi = ia. ρµ 3 With that, we can substitute ρµ(xi) ± a∂xρµ(xi) + 1 xρµ(xi) + O(a3) for ρµ i±1 and the currents are 2 a2∂2 J T(x) = ρT(x) − f (x) J S(x) = −a∂xρS(x) − 1 2 a2∂2 xρS(x) + O(a3) where f (x) is the continuous version of fi. In the steady-state Eq. (S11) translates to 0 = −a∂x(J T + J S) + 1 2 a2∂2 x(J T + J S) + O(a3) = = −a∂x(ρT(x) − f (x)) + O(a2) to first order in a and thus, we have ρT(x) − f (x) = const to lowest order in a. Using Eq. (S7) or (S8) we conclude that const = 0 or that the combined current J T i must be constant and zero everywhere. This corresponds to the fact that the system is closed at the tip so that in steady state on average there is no influx into the system from the base. Therefore, i + J S ρT(x) ≡ f (x) or, equivalently, i = fi =(cid:10)nT i nT i+1 (cid:11) = Prob{nT Prob{nT i } = ρT Prob(cid:0)nT holds to lowest order in a where we defined Prob{nT i = 1(cid:1) analogously to before. This implies that either i+1 = 1nT i , nT i , nT i+1} = Prob{nT i+1} = Prob(cid:0)nT i = 1, nT i+1nT i }Prob{nT i } i+1 = 1(cid:1) and Prob{nT i+1nT i } = or i = Prob{nT ρT i } = 0 (S14) Prob{nT i+1nT i } = 1. The latter implies that, whenever site i is occupied, site i + 1 is occupied with probability 1 as well. Hence, if there is some site j occupied at some time, any site i > j is occupied at that time as well. It follows that to lowest order in a the density suddenly jumps from zero (S14) to maximal density ρT i = 1. (S15) This means that to lowest order the steady-state TASEP profile is given by a step function separating a region of 0 density on the left (towards the base) from a region of density 1 on the right (towards the tip) and, in particular, there is no maximal current phase as it occurs for TASEP alone. Instead, the generic TASEP profile is given by a domain wall and we make the following fully-localised-DW ansatz: (cid:40) ρT(x) = for x ∈ [0, z[ for x ∈ ]z, 1] 0 1 (S16) where z ∈ [0, 1] is the position of the domain wall (step) that we will determine later. We will now use this ansatz to determine the steady-state density profile for the SSEP lane depending on z. For this purpose, let us go back to Eq. (S2) that is given by 0 = a2∂2 xρS(x) + Ωa(ρT(x) − ρS(x)) in the continuum limit in the steady state. This equation can be solved in the two regions x < z and x > z as A1 cosh (cid:16)(cid:113) Ω (cid:17) (cid:16)(cid:113) Ω a x (cid:17) + A2 sinh (cid:16)(cid:113) Ω (cid:17) (cid:16)(cid:113) Ω a x (cid:17) ρS(x) = for x ∈ [0, z[ for x ∈ ]z, 1]. The constants A1, A2, B1 and B2 can be determined from the boundary conditions: Eq. (S5) yields β(cid:0)ρT 1 + B1 cosh + B2 sinh a x a x L −(cid:10)nT LnS L 0 to lowest order in a for model A, and assuming that nT supported by the simulations) we conclude that L and nS L are uncorrelated (mean-field assumption that is 4 (cid:11)(cid:1) = for model A, unless ρT L = 0 and so z = 1. For model B, we can use Eq. (S6) and find ρS L = 1 (cid:0)1 − ρS(1)(cid:1) , ∂xρS(1) = Ω  again unless z = 1. This latter case needs to be treated separately. This can be done for instance by regarding the "domain wall" as a boundary layer with slope (ρS(1)right − ρS(1)left)/a where ρS(1)right is the density at the very last site L and ρS(1)left is the limit x → 1 of the low-density phase. Basically, this case can be understood as the limit where the domain wall is shifted to the right out of the system. As long as  ∈ O(a0), the generic density profile is given by a domain wall separating density 0 on the left from density 1 on the right, that is ρT ∈ {0, 1} holds to lowest order in a. In the limiting cases ρT ≡ 1 the density reaches the value of 1 in only a few lattice sites from the base, whereas in the case ρT ≡ 0 the density has a very small spike (boundary layer) only at the tip. The details of this calculation are, however, out of the scope of this letter and we will carry on with the treatment of the case z (cid:54)= 1. i+1 − ρS steady-state density profile and the derivative thereof is continuous (it can be seen from Eq. (S2) that(cid:0)ρS (cid:0)ρS For both models we have Eq. (S13) for the boundary condition at the base. Requiring that both the diffusive i − ρS i−1 (cid:1) = O(a) holds), the following expression for the SSEP profiles can be derived: (cid:1) − i ρS(x) = for model A and for model B where (cid:40) l σ cosh(cid:0) x 1 −(cid:0)cosh(cid:0) z (cid:40) σ cosh(cid:0) x 1 −(cid:0)cosh(cid:0) z l l l ρS(x) = (cid:1) l l l l l l l l (cid:1)(cid:1) sinh(cid:0) x (cid:1) − σ(cid:1) − sinh(cid:0) z (cid:1) +(cid:0)coth(cid:0) 1 (cid:1)(cid:0)cosh(cid:0) z (cid:1)(cid:1) (cid:1) − coth(cid:0) 1 (cid:1) sinh(cid:0) x (cid:1) − σ(cid:1)(cid:0)cosh(cid:0) x (cid:1)(cid:1) sinh(cid:0) x (cid:1) − σ(cid:1) − sinh(cid:0) z (cid:1) +(cid:0)γ(cid:0)cosh(cid:0) z (cid:1) (cid:1)(cid:1) (cid:1) − γ sinh(cid:0) x (cid:1) − σ(cid:1)(cid:0)cosh(cid:0) x (cid:114) (cid:114) (cid:114) a (cid:1) + a sinh(cid:0) 1 l cosh(cid:0) 1 (cid:1) + a l sinh(cid:0) 1 cosh(cid:0) 1 (cid:1) (cid:1) ≈ tanh (cid:18) 1 a2 ω (cid:19) D ω γ = l = = = Ω l . l l l l l l l l with the diffusion constant D = a2 and for x ∈ [0, z[ for x ∈ ]z, 1] for x ∈ [0, z[ for x ∈ ]z, 1] (S17) (S18) To illustrate the differences between the two models, Fig. S1 shows typical density profiles on the SSEP lane for both cases. As one can see, the density profiles differ significantly between the two models. For model A, the density generically increases considerably from the protrusion base with density σ towards the tip with density 1 (unless z = 1). In contrast, the density for model B at the tip reaches a value that is only slightly larger than the value at the base. The fact that for model A the cytoplasmic density at the tip reaches a value very close to 1 is also the reason why for model A, the exclusion in the cytoplasm has a much higher influence on the residence time at the tip compared to model B where the cytoplasm is occupied quite homogeneously and where exclusion at the tip is not more important than in the bulk. Using Eqs. (S17) and (S18) we now have expressions for the steady-state density profiles that only depend on one quantity, namely z, the position of the domain wall. Using the mass-balance equations [Eq. (2) and the corresponding one for model B], this enables us to find an analytic formula for z. fi − fi−1 = ρT i(cid:88) Hence, we can write i(cid:88) (cid:0)ρT Using Eq. (S4) we can rewrite f0 = ρT + ω(cid:0)ρT (fj − fj−1) = f0 + fi = f0 + j=1 j=1 0 − ρS 0 j − ρT i i − ρS i − ρT i−1 + ω(cid:0)ρT j−1 + ω(cid:0)ρT (cid:1) and, thus, i(cid:88) (cid:0)ρT j − ρS j − ρS (cid:1) . (cid:1)(cid:1) = f0 + ρT (cid:1) . j j i − ρT 0 + ω (cid:0)ρT j − ρS j (cid:1) . i(cid:88) j=1 5 FIG. S1. Steady-state SSEP density profiles are shown examplarily for model A (left panel) with L = 50, β = 0.2, Ω = 0.001, α = 0.1 and  = 0.025 and for model B (right panel) with L = 50, Ω = 0.001, α = 0.2 and  = 0.3. The simulation results (filled red circles) agree well with the theoretical prediction (dotted green curve) according to Eq. (S17) and (S18), respectively. For model A, the density is increasing towards the tip where it has a high slope and reaches a value of 1. For model B, the density is more homogeneous and has a small slope at the tip. POSITION OF THE DOMAIN WALL In order to determine the position of the domain wall, we go back to Eq. (S1). In the steady state this reduces to fi = ρT i + ω (S19) nearest-neighbour correlator fi =(cid:10)nT This is the moment-identity [Eq. (1)]. i nT i+1 j=0 (cid:11) and the average densities. We will use it later to predict the covariances It is an important result for us as it gives an exact relation between the and the TASEP current in the system. This will allow us to make a sharp distinction between the actual TASEP current J T i and the mean-field current J T MF,i = ρT i i+1 From the moment-identity, Eq. (S19), we can also easily deduce the mass-balance equations [Eq. (2) and the corresponding one for model B] by using the boundary conditions at the tip [Eq. (S5) for model A and (S6) for model B] and the moment-identity for i = L − 1: (cid:0)1 − ρT (cid:1). ω model A : i − ρS (cid:0)ρT (cid:0)ρT L −(cid:10)nT L−1(cid:88) L(cid:88) equations in the continuum limit: Using that(cid:82) 1 0 dx ρT(x) = 1 − z and that(cid:80)L (cid:90) 1 (cid:1) = −β(cid:0)ρT (cid:1) = 0. we find to lowest order in a: i − ρS model B : (cid:18) (cid:19) i=0 i=0 ω i i (cid:11)(cid:1) LnS L In order to determine z we will now insert the domain wall ansatz, Eq. (S16), for the TASEP together with the respective SSEP profile into the respective mass-balance equation. For this purpose, we need to write the mass-balance 0 dx ρµ(x) 0 di ρµ(ia) = 1 i=0 ρµ a i ≈(cid:82) L (cid:82) 1 = −β(cid:0)ρT(1)(cid:0)1 − ρS(1)(cid:1)(cid:1) = −a∂xρS(1) model A : Ω model B : 1 − z − 1 − z − (cid:90) 1 0 dx ρS(x) 0 dx ρS(x) = 0 where for model A the mean-field assumption(cid:10)nT explicit densities (S17) or (S18) into these expressions, integrating and cancelling common factors results in LnS L LρS L was used again, together with Eq. (S5). Inserting the (cid:11) ≈ ρT (cid:18) 1 (cid:18) 1 l (cid:19) (cid:19) model A : model B : σ cosh = cosh σ sinh = sinh l l (cid:18) 1 − z (cid:19) (cid:19) (cid:18) 1 − z (cid:19)(cid:19) (cid:18) 1 (cid:18) 1 (cid:19)(cid:19) l l . (cid:18) (cid:18) model A : z = 1 − l cosh −1 σ cosh model B : z = 1 − l sinh −1 σ sinh . l 6 (S20) (S21) (S22) (S23) Those directly lead to the expressions for the domain wall given in the main text: or l = 1/ cosh We see that z is shifted towards the tip (base) for smaller (higher) values of σ := α that for model A the limit z (cid:37) 1 is reached for finite values of the parameters σ, ω and  when σ cosh(cid:0) 1 the fact that σ cosh(cid:0) 1 −1(cid:0)σ sinh(cid:0) 1 parameter regimes where there is no domain wall and where to lowest order the density is zero everywhere. For model B, this case only occurs in the limit where the cytoplasmic density at the base is zero. Even for large l we have z = 1 − l sinh (cid:1) lacks a real solution for z if σ cosh(cid:0) 1 −1(cid:0) σ (cid:1) = 1 (cid:1) < 1. That is, for model A there are −1(1/σ), whereas for model B, z (cid:37) 1 is only possible in the limit σ → 0. This is also reflected in (cid:1) ≈ 1 − σ and thus, z < 1 unless σ = 0. (cid:1) = cosh(cid:0) 1−z (cid:1)(cid:1) ≈ 1 − l sinh It is interesting to note α+ . l l l l l l REFINED MEAN-FIELD TASEP DENSITY PROFILE Certainly, the TASEP profile is not just given by a plain step function but has a smooth form that separates the two density regions via an intermediate region where the density increases strongly but whose width is finite. This is due to fluctuations in the stochastic system that soften this transition. Our first approach to resolve this finite width relies on an idea similar to the method used in [39], namely to split the system into a part around the domain wall and the parts further away from it. By this, only parts of the terms contribute, respectively, and assuming a steep domain wall for the TASEP one can neglect the terms stemming from the exchange with the diffusive lane in this narrow region: Reconsidering equation (S1) we realize that in the steady state it can be approximated as (cid:0)ρT(cid:0)ρT − 1(cid:1)(cid:1) + 0 = a∂x xρT + Ωa(cid:0)ρS − ρT(cid:1) 1 2 a2∂2 i nT where we used the continuum limit and a mean-field assumption for the nearest-neighbours on the TASEP: fi = i+1. If we want to investigate the vicinity of the domain wall x ≈ z this can be done by considering x := (x − z)/a that is very large away from the domain wall. Using this coordinate system the above equation looks like follows: i ρT i+1 (cid:10)nT (cid:11) ≈ ρT Dropping the term of order O(a), we end up with xρT(x) + Ωa(cid:0)ρS(x) − ρT(x)(cid:1) . ∂2 0 = ∂x (cid:0)ρT(x)(cid:0)ρT(x) − 1(cid:1)(cid:1) + ρT(x)(cid:0)ρT(x) − 1(cid:1) + 1 2 where const can be estimated from the boundary conditions for x → ±∞: we have ρT(x = 0) = 1 − ρT(x = 1) = 0 and ∂xρT(x)(cid:12)(cid:12)x=0/1 ≈ 0, and hence, using x → ±∞ we find that const ≈ 0. As a result, after integrating ρT(x)(cid:0)ρT(x) − 1(cid:1) + 1 2 ∂xρT(x) = 0 with respect to x, using that ρT(z) = 1 2 , and rescaling we find 1 2 ∂xρT(x) = const (cid:20) (cid:18) x − z (cid:19)(cid:21) a ρT(x) ≈ 1 2 1 + tanh for the refined mean-field TASEP density profile. Contrary to the step profile, Eq. (S16), it exhibits a finite width that scales like a or inversely proportionally to L. But as one can infer from Fig. 2 it still underestimates the actual width due to fluctuations that are ignored by the mean-field assumption for the nearest-neighbours on the TASEP. DOMAIN WALL THEORY 7 This is why we also pursue another approach to refine our prediction for the domain wall, namely by going beyond mean-field theory and treating the domain wall as a random walker that moves in a non-uniform potential with reflecting boundaries. This technique has been introduced in Ref. [48] for the TASEP and, more generally, for shock waves in Ref. [49]. It has been applied to TASEP-LK in Refs. [18, 19] using non-uniform hopping rates. We follow their ideas and describe the domain wall by site-dependent hopping rates wl,i = JLD,i HD,i − ρT ρT LD,i wr,i = JHD,i HD,i − ρT ρT LD,i (S24) to the left and right from site i, respectively. Here, JLD,i (JHD,i) denotes the steady-state TASEP current at site i under the assumption that site i is in the low-density (high-density) phase, i.e. on the left (right) of the domain wall. ρT LD,i and ρT HD,i are the respective TASEP densities in the low- and high-density phase. The heuristic reason why the hopping rates have the above form, Eq. (S24), is that the low-density current JLD,i should just correspond to the current arriving at the left of the domain wall and causing the domain wall to move one step to the left. Thereby, at site i the density increases by ρT LD,i at rate wl,i. Similarly, the high-density current JHD,i should correspond to the current leaving the region at the right of the domain wall, and causing the domain wall to move one step to the left. This happens at rate wr,i and the density decreases by ρT HD,i − ρT HD,i − ρT LD,i. From the detailed balance condition ps,iwr,i = ps,i+1wl,i+1 it is easy to see that the stationary distribution of the shock position ps,i satisfies ps,i ∝ 1 wl,i exp − i−1(cid:88) j=1 (cid:19) . (cid:18) wl,j wr,j ln We will show next that wl,iz = wr,iz holds for iz being the discrete domain wall position (iz = z/a): From Eq. (S1) as well as from the definition of the local TASEP current, Eq. (S9), we know that J T we have i − ρT i = J T i i−1 + ω(cid:0)ρS (S25) (cid:1). Hence, (cid:1) , i(cid:88) (cid:0)ρS j − ρT L−1(cid:88) (cid:0)ρS j=1 j j − ρT j j=i+1 or (cid:1) . J T i = J T 0 + ω J T i = J T L−1 − ω As mentioned after introducing JLD/HD,i in Eq. (S24), JLD/HD,i denotes the TASEP current at site i assuming that this site is in the low-density (high-density) region. That is, we can calculate JLD,i (JHD,i) by using the above iterations starting from i = 0 (i = L − 1) and assuming that between site 0 (site L − 1) and site i the densities are given in the low-density (high-density) phase: JLD,i = JLD,0 + ω j=1 JHD,i = JHD,L−1 − ω i(cid:88) (cid:1) , LD,j (cid:0)ρS LD,j − ρT L−1(cid:88) (cid:0)ρS HD,j − ρT HD,j j=i+1 or (cid:1) . (S26) (S27) We want to compare those two currents right at the domain wall i = iz. For this purpose, we use that JLD,j = J T j for j < iz and JHD,j = J T for j > iz since we chose z in such a way that the densities to the left (right) of the fixed j domain wall iz are in the low-density (high-density) phase. As a result, L−1(cid:88) L −(cid:10)nT (cid:1) − J T (cid:0)ρS (cid:11)(cid:1) for model A and J T HD,j − ρT 0 − ω iz(cid:88) (cid:0)ρS L−1 = ω(cid:0)ρT HD,j j=1 j=iz+1 j − ρT j L − ρS L L−1(cid:88) (cid:0)ρS (cid:1) = 0 (cid:1) = J T (cid:1) for model B and, hence, the above L−1 − ω j − ρT j=0 j L−1 = β(cid:0)ρT JHD,iz − JLD,iz = JHD,L−1 − ω for both models: J T holds due to the respective mass-balance equation. It follows that LnS L wl,iz = wr,iz (S28) (cid:16) wl,iz (cid:17) (cid:16) wl,j (cid:17) > 0 for j > iz such that the domain and, thus, ln wall preferably walks towards the (fixed) domain wall position iz than away from it. Therefore, the exponential in Eq. (S25) has a maximum at i = iz and taking the continuous version of this equation = 0. Similarly, we have ln < 0 for j < iz and ln wr,iz wr,j ps(x) ∝ 1 exp wl(x) dx(cid:48) ln (cid:20) ≈ 1 wl(z) exp − 1 a wr,j (cid:17) (cid:16) wl,j (cid:18) wl(x(cid:48)) (cid:19)(cid:21) (cid:18) wl(x(cid:48)) (cid:19)(cid:21) (cid:20) wr(x(cid:48)) wr(x(cid:48)) exp (cid:20) − 1 2a (cid:90) x 0 dx(cid:48) ln (cid:20) − 1 a (cid:90) z 0 (cid:21) we can use the method of steepest descent to approximate (cid:20) 1 wl(x) exp − 1 a (cid:19)(cid:21) (cid:18) wl(x(cid:48)) wr(x(cid:48)) 0 dx(cid:48) ln (cid:90) x (cid:20) (cid:20) − 1 2a − (x − z)2 W (z)2 (x − z)2 J(cid:48) (cid:21) = exp 8 (S29) (cid:21) (cid:19)(cid:12)(cid:12)(cid:12)(cid:12)x=z (cid:18) wl(x) (cid:21) wr(x) − 1 2a (x − z)2∂x ln where wl(x) and wr(x) is the continuous version of wl,i and wr,i, respectively. Since wl(x) (x − z)2 J(cid:48) LD(z)JHD(z) − J(cid:48) HD(z)JLD(z) ps(x) ∝ exp = exp JHD(x) , we conclude wr(x) = JLD(x) LD(z) − J(cid:48) JLD(z) HD(z) = (S30) JLD(z)JHD(z) where we used that JLD(z) = JHD(z) [Eq. (S28)] and defined the width W (z) by To find a formula for ρT(x) we realize that W −2(z) = HD(z) . LD(z) − J(cid:48) J(cid:48) 2aJLD(z) (cid:90) 1 (cid:90) x 0 (cid:90) x 0 dx ps(x) dx ps(x) ≈ x ρT(x) = ρT HD(x) dx ps(x) + ρT LD(x) HD(x) ≈ 1 and ρT since as along as the shock position is left (right) of x there is a high-density (low-density) region at x and since we can approximate ρT (cid:17) (cid:17) where erf is the error function. Here, we used that we need to normalize ps(x) such that(cid:82) 1 (cid:82) x LD(x) ≈ 0 to lowest order. Using Eq. (S30) we can determine ρT(x) as (cid:82) 1 (cid:17) (cid:16)− (x−z)2 (cid:17) = (cid:16)− (x−z)2 (cid:16) x−z (cid:16) 1−z (cid:16) z (cid:16) z ρT(x) ≈ 0 dx exp 0 dx e (cid:17) (cid:17) (S31) + erf + erf W (z)2 W (z)2 W (z) W (z) W (z) W (z) erf erf 0 dx ps(x) = 1. In order to use this formula we still need to determine the width W (z) more concretely. For this purpose, we must make some assumption on how we treat the diffusive (SSEP) lane when considering the low- and high-density currents. We are not aware of previous attempts that apply domain wall theory in case where a TASEP is coupled to another lattice that is occupied stochastically as well. Then the attachment and detachment rates for the TASEP not only depend on the TASEP occupancy but also on the occupancy on the other lattice. The difficulty with this situation is that it is a priori not clear if one should assume that the occupancy on the coupled lattice is fixed, that is, does not change very much if the domain wall is shifted towards the left or right, or if one should look at the momentary steady-state density corresponding to a certain position of the domain wall. For simplicity, we chose the first ansatz that seems to be working well. That means, we use Eqs. (S17) and (S18) for the SSEP density profile with the calculated fixed position of the domain wall z given by Eqs. (S20) and (S21) and we assume that this SSEP density does not change considerably when the domain wall is shifted shortly to some other position (cid:54)= z, i.e. LD(x) = ρS ρS With that in mind, using Eqs. (S26) and (S27), we find that HD(x) = ρS(x). and obtain model A : model B : JHD(x) = −Ω JHD(x) = −Ω JLD(x) = Ω 0 dx ρS(x) (cid:90) x (cid:90) 1 dx (cid:0)ρS(x) − 1(cid:1) − a∂xρS(1) (cid:90) 1 dx (cid:0)ρS(x) − 1(cid:1) x x 9 FIG. S2. Comparison between the mean-field current (left panel) and the actual current (right panel) for model A with L = 50, β = 0.2, Ω = 0.001, α = 0.1 and  = 0.025. The mean-field current (filled blue (grey) circles) is obtained from the average densities in a simulation. On the right, we show the actual currents as measured directly in a simulation (filled red (dark grey) i − fi circles), as obtained from the average densities and the nearest neighbour correlations in the simulation using J T (filled orange (grey) circles) and as predicted by the theory from Eq. (S32) (dotted green (light grey) curve). i = ρT where we used that ρT LD(x) ≈ 0 and ρT HD(x) ≈ 1. As a result, J(cid:48) LD(x) = ΩρS(x) and J(cid:48) HD(x) = Ω(cid:0)ρS(x) − 1(cid:1) so that (cid:115) (cid:90) z 0 (cid:114) (cid:16) z (cid:17) l W (z) = 2a dx ρS(x) = 2σal sinh where the last equality results from integrating ρS and then using the defining equation for z. Interestingly, the result in this form agrees for both models. the limit a → 0 it holds that l → 0 (if we keep  constant) so that we can approximate arccosh(cid:0)σ cosh(cid:0) 1 and, similarly, arcsinh(cid:0)σ sinh(cid:0) 1 We observe that the width W (z) increases with increasing z or decreasing distance from the tip. Furthermore, in l +ln (σ) l + ln (σ). Thus, z ≈ −l ln (σ) for both models in the limit a → 0 [Eqs. (S22), (cid:1)(cid:1) ≈ 1 (cid:1)(cid:1) ≈ 1 l l (S23)]. With that we find (cid:0)1 − σ2(cid:1) 1 2 4  1 W (z) ≈ a 3 Ω 1 4 4 sum(cid:80)L in the limit a → 0 and so the width of the domain wall decreases with decreasing a or increasing number of sites L. Note that in order to calculate the position of the domain wall z by the mass-balance equation, we can use the step function instead of the refined profiles. This is due to the fact that there the errors more or less cancel since only the i over the densities on the left and on the right of the domain wall enters. However, this approximation gets worse the closer the calculated z is to 0 or 1 since then, the errors are not symmetric anymore. As a result, also the approximation, Eq. (S31), deteriorates. Certainly, if there is no solution for z, e.g. for some parameters in model A, we can not use this ansatz either. Furthermore, if  is too large, we expect our approximation that ρS(x) does not change when the domain wall is shifted to deteriorate as well. i=0 ρT COVARIANCES AND CURRENTS Another point we want to emphasize is that in the system the nearest-neighbour correlations of the TASEP sig- nificantly modify the TASEP current in the sense that the mean-field prediction for it overestimates the current by orders of magnitude. The reason lies in the covariances that reach very high values of around 0.2 in the region of the domain wall (see for example Fig. 2). As we have shown above in Eq. (S19) the nearest-neighbour correlator fi =(cid:10)nT i nT i+1 (cid:11) is given by fi = ρT i + ω(cid:80)i j=0 (cid:0)ρT (cid:1). Therefore, the covariances are given by j − ρS i(cid:88) i(cid:88) j j − ρS (ρT j ) = J T MF,i + ω j − ρS (ρT j ). Covi = ρT i (1 − ρT i+1) + ω Due to the first term, the mean-field current, which is large (≈ 0.2) in the region of the domain wall, also the covariances are non-zero and large there. However, since the mean-field current gives the main contribution to the j=0 j=0 10 left(x) on the SSEP towards the base (left) (filled orange (light grey) squares) and J S FIG. S3. Illustration of the different currents in model A with L = 200, β = 0.2, Ω = 0.1, α = 0.1 and  = 1.55. This choice of parameters corresponds to a domain wall position of z ≈ 0.88 so that the TASEP occupancy is concentrated on a small region around the tip. The left panel shows a comparison of the current J T(x) on the TASEP (filled red (dark grey) circles) and the currents J S right(x) towards the tip (right) (filled green (grey) diamonds). The TASEP current corresponds to the net SSEP current towards the base, that is the difference between the SSEP current towards the base and towards the tip. Both SSEP currents, however, are much larger than the TASEP current suggesting that the motors are mainly transported by the cytoplasm (SSEP) rather than by the filament (TASEP). The right panel shows both the attachment current Jon(x) (filled purple (dark grey) downward facing triangles) and the detachment current Joff (x) (light blue (light grey) upward facing triangles). Both currents are enhanced considerably towards the tip region. As a result, motor exchange happens primarily around the tip region. Furthermore, the maximum of the attachment current is located a little further away from the tip than the maximum of the detachment current so that typically motors attach to the filament, walk a short distance on the filament, detach near the tip and can then diffuse in the cytoplasm back to the cell body or reattach again. Note that the detachment current is particularly high at the last site due to the higher detachment rate at the tip. All currents are measured directly from the simulation that is counting the number of jumps per time. covariances, this implies that the actual current J T i = ρT i − fi = ω i(cid:88) j=0 (cid:0)ρS j − ρT j (cid:1) = J T MF,i − Covi (S32) is much smaller than predicted from a mean-field theory. To illustrate this further, in Fig. S2 we show a comparison of the mean-field current on the left as obtained from simulations using the average densities, and of the actual current on the right as obtained directly from the simulation (filled red (dark grey) circles), from the average densities and nearest-neighbour correlations in the simulation (filled orange (grey) circles) and from Eq. (S32) (dotted green (light grey) curve). Certainly, this discrepancy scales with Ω (and the other parameters) and the case shown here is extreme as Ω = 0.001 was chosen. However, the fact that both currents differ by orders of magnitude is robust and also occurs for Ω being of the order of the other jump rates [Fig. S3]. This shows that a mean-field description fails to capture essential properties of our model system. Furthermore, from the fact that the covariances are non-zero everywhere Covi ≥ 0 and that we can rewrite i }. And from using a similar argument and only interchanging the we can conclude that Prob{nT i+1}. This tells us that, in particular from the domain wall roles of i and i + 1 we find Prob{nT region onwards where covariances are high, the motors preferentially cluster and hinder each other effectively so that the mean time a particle spends at a certain site is increased considerably with respect to freely moving motors. Eventually, this leads to the substantial difference between the actual current and the mean-field current. i+1} ≥ Prob{nT i+1nT i } ≥ Prob{nT Potentially, those results are also important from a biological point of view. In case of an enhanced detachment rate at the filament's tip (model A), tip localisation is facilitated [Fig. 3] and large traffic jams can be avoided. In this case there is only high filament density in a small region around the tip so that motors spend more time in the tip region than in the other part of the filament, a feature that might be favoured biologically as then the motors or their cargo might have more time to perform the necessary tasks at the tip. Furthermore, transport to the tip might be strongly promoted by diffusion in the cytoplasm (SSEP) whose currents in both the direction of the tip Covi =(cid:10)nT (cid:11) − ρT i nT i+1 = Prob{nT i nT i+1 = Prob{nT i ρT i , nT i nT i+1}Prob{nT i+1} − Prob{nT i+1} − Prob{nT i }Prob{nT i+1} i }Prob{nT i+1} = 11 FIG. S4. Illustration of the generalised model with several (here: 4) lanes for diffusion: the three-dimensional view is shown on the left-hand side, the profile is illustrated on the right-hand side. As before the dynamics on the filament (red) are given by a TASEP with jump rate ν = 1 but now the dynamics in the cytoplasm is modelled by several lanes for diffusion (blue), each with diffusive rate  and respecting the exclusion. The lanes for diffusion are arranged in a cylinder-like fashion around the TASEP and each can interact with the TASEP by attachment/detachment processes at rate ω respecting the exclusion. Apart from the diffusion along the cylinder axis there is also lateral diffusion between neighbouring lanes of diffusion. This diffusion happens at rate lat and again respects the exclusion. At the base (not shown) there is influx at rate α into every lane for diffusion and outflux at the diffusive rate . and of the base significantly exceed the filament (TASEP) current [Fig. S3]. Certainly, in case of our original model where exclusion in the cytoplasm occurs at occupancy 1, the motors cannot really bypass each other, so that the motors do not circulate very often. This, however, should be greatly enhanced in the generalised model with carrying capacity Nmax > 1 (see next chapter) or in a real biological system where the motors can overtake each other in the cytoplasm. The fact that the filament current is strongly suppressed by excluded volume effects might also be beneficial from a biological point of view as every motor step on the filament consumes ATP contrary to diffusion in the cytoplasm. Thus, transport of motors by the cytoplasm rather than by the filament might be advantageous energetically. Moreover, both the attachment and detachment current Jon and Joff are mainly restricted to the tip area [Fig. S3] so that motor exchange between the filament and the cytoplasm occurs primarily near the tip where the cargo is used. Taken together, tip localisation and the suppression of the filament current might be beneficial from a biological point of view as then energy consumption is low and motors could be efficiently transported to the tip by the cytoplasm. Near the tip they attach to the filament, have an enhanced residence time on the filament due to steric hindrance between the motors and then detach at the tip back into the cytoplasm. CYLINDRICAL GEOMETRY WITH SEVERAL LANES FOR DIFFUSION Finally, we want to deal with a generalisation of our model where instead of one lane for diffusion we have several lanes for diffusion arranged on a cylinder around the filament. For an illustration of the case with Ndiff = 4 lanes for diffusion please refer to Fig. S4. The three-dimensional view is shown on the left-hand side, the profile is illustrated on the right-hand side. As before the dynamics on the filament (red) are given by a TASEP lane with jump rate ν = 1 but now the dynamics in the cytoplasm is modelled by several lanes for diffusion (blue), each with diffusive rate  and respecting the exclusion. The lanes for diffusion are arranged in a cylinder-like fashion around the TASEP lane, and each can interact with the TASEP lane by attachment/detachment processes at rate ω respecting the exclusion. Apart from the diffusion along the cylinder axis there is also lateral diffusion between neighbouring lanes of diffusion. This diffusion happens at rate lat and again respects the exclusion. At the base (not shown) there is influx at rate α into every lane for diffusion and outflux at the diffusive rate . In the following, we will show that we can reduce the steady-state behaviour of this more elaborate model to the steady-state behaviour of our model by scaling the parameters for diffusion and attachment/detachment by the number of lanes for diffusion Ndiff : ω → ωNdiff ,  → Ndiff . Here, we only consider the case where we have the same attachment and detachment kinetics at the tip than in the bulk (model B), but an analogous generalisation νωnTinS,1inS,4inS,3inS,2i (cid:16) ρS,m+1 L + ρS,m−1 L − 2ρS,m L (S35) (cid:17) . + lat Due to the exclusion property on both the TASEP as well as on the lanes for diffusion the state space of the system is finite. Furthermore, it is an irreducible continuous-time Markov process so there exists a unique steady-state. This is important since then the cylindrical symmetry of the system must be reflected in this steady-state. This is why i (cid:105) holds for all m, n. As a result, we can assume that in the steady-state the terms proportional to lat drop out in Eqs. (S33)-(S35) and we can define the total occupancy at site i in the cytoplasm i (cid:105) = (cid:104)nS,n and (cid:104)nS,m i nT i nT nS,m i nS,n i (cid:69) = 0 1 − ρS,m ρS,m Ndiff(cid:88) (cid:17) m=1 ω + ω L )(cid:105)(cid:17) L)(cid:105)(cid:17) L(1 − nS,m L (1 − nT (cid:16)(cid:104)nS,m (cid:16)(cid:104)nT (cid:68) L (1 − nT L)(cid:105) − (cid:104)nT L )(cid:105) − (cid:104)nS,m L(1 − nS,m (cid:68) (cid:69) i ≡ Ndiff(cid:88) nS m=1 nS,m i 12 (S36) (S37) (S38) and argument can be done for model A or different rates for attachment and detachment. Let us denote by nT i the occupancy at site i of the TASEP (as before) and by nS,m the occupancy at site i of the m = 1, . . . , Ndiff -th lane for diffusion. We start by m = 1 at an arbitrary lane for diffusion and then consecutively number the lanes for diffusion in clockwise order. Since the lanes are arranged on a cylinder we have periodic boundary conditions and identify nS,Ndiff +1 . With this convention we have the following bulk master equations, written straight away in terms i of the averages over occupancies ρµ,m ≡ nS,1 (cid:11): i i i nT i+1 i (cid:105) and fi =(cid:10)nT (cid:16)(cid:104)nS,m (cid:16)(cid:104)nT i (1 − nS,m (1 − nT i = (cid:104)nµ,m Ndiff(cid:88) (cid:17) + ω m=1 ω i i )(cid:105)(cid:17) i )(cid:105)(cid:17) i )(cid:105) − (cid:104)nT )(cid:105) − (cid:104)nS,m i (1 − nS,m (1 − nT i i (cid:16) ρS,m+1 i + ρS,m−1 i − 2ρS,m i (cid:17) . (S33) + lat )(cid:105)(cid:17) (1 − nT 0 (1 − nS,m 0 0 )(cid:105) − (cid:104)nT (cid:17) (cid:16)(cid:104)nT + ω 0 (1 − nS,m 0 )(cid:105) − (cid:104)nS,m 0 0 )(cid:105)(cid:17) (1 − nT + lat (cid:16) (S34) (cid:17) ρS,m+1 0 + ρS,m−1 0 − 2ρS,m 0 At the base we find i ∂tρT ∂tρS,m i =  i = ρT i+1 + ρS,m ρS,m i−1 − fi−1 − ρT (cid:16) i + fi + i−1 − 2ρS,m Ndiff(cid:88) 0 = −ρT 0 = α − (α + )ρS,m (cid:16)(cid:104)nS,m (cid:16) 0 + f0 + 0 +  m=1 ω 0 ∂tρT ∂tρS,m and at the tip ∂tρT L = ρT ∂tρS,m L =  L−1 − fL−1 + (cid:16) L−1 − ρS,m ρS,m L and the corresponding average ρS together. Using this quantity we can rewrite Eqs. (S33)-(S35) in steady-state as follows: i as the sum of the (average) occupancies at sites i of all the lanes for diffusion taken for the bulk master equation, for the base and i + fi + ω(cid:0)(cid:104)nS (cid:1) + ω(cid:0)(cid:104)nT i−1 − 2ρS i i )(cid:105) − (cid:104)nT i (1 − nT i (Ndiff − nS i (Ndiff − nS i (1 − nT i )(cid:105) − (cid:104)nS i )(cid:105)(cid:1) i )(cid:105)(cid:1) i−1 − fi−1 − ρT 0 = ρT i+1 + ρS 0 = (cid:0)ρS 0 + f0 + ω(cid:0)(cid:104)nS 0 = −ρT 0 = αNdiff − (α + )ρS 0 0 )(cid:105) − (cid:104)nT 1 − ρS 0(1 − nT 0 + (cid:0)ρS L−1 − fL−1 + ω(cid:0)(cid:104)nS (cid:1) + ω(cid:0)(cid:104)nT L−1 − ρS L 0 = ρT 0 = (cid:0)ρS 0)(cid:105)(cid:1) 0 (Ndiff − nS (cid:1) + ω(cid:0)(cid:104)nT 0 (Ndiff − nS 0)(cid:105) − (cid:104)nS 0(1 − nT 0 )(cid:105)(cid:1) L)(cid:105) − (cid:104)nT L(1 − nT L(Ndiff − nS L(Ndiff − nS L(1 − nT L)(cid:105) − (cid:104)nS L)(cid:105)(cid:1) L)(cid:105)(cid:1) for the tip. From these equations it becomes apparent that the cylindrical system with Ndiff lanes for diffusion is equivalent to the above mentioned generalisation of our model where exclusion on the filament does not happen at occupancy of 1 but at a maximal occupancy or carrying capacity of Nmax = Ndiff that is reflected in the term Ndiff − nS i . To show that the steady-state behaviour of those two generalisations of our model are cast by the steady-state behaviour of our model, we next introduce the quantity nS i = 1 Ndiff nS i = 1 Ndiff nS,m i Ndiff(cid:88) m=1 13 and the corresponding average occupancy ρS i at site i of one lane for diffusion. In terms of nS i we find (cid:0)ρS i−1 − fi−1 − ρT i+1 + ρS 0 = ρT 0 = Ndiff (cid:0)ρS (cid:1) + ωNdiff (cid:1) i − ρS (cid:0)ρT i i − ρT i (cid:1) for the bulk and i i + fi + ωNdiff i−1 − 2ρS (cid:1) (cid:0)ρS (cid:1) (cid:0)ρT L − ρS 0 − ρT 0 + Ndiff L − ρT (cid:0)ρS (cid:0)ρS (cid:1) + ωNdiff L 0 0 1 − ρS (cid:1) i (1 − nS L 0 + f0 + ωNdiff 0 = −ρT 0 = αNdiff − (α + )Ndiff ρS L−1 − fL−1 + ωNdiff 0 = ρT (cid:0)ρS (S39) (S40) (cid:1) (S41) (cid:1) + ωNdiff (cid:0)ρT 0 − ρS 0 L 0 = Ndiff L−1 − ρS for the base and the tip where we used that (cid:104)nS In summary, if we replace ω → ωNdiff ,  → Ndiff and α → αNdiff in our model we can deduce from the steady-state behaviour of our model the steady-state behaviour of these generalised models. Certainly, to stay within the scope of our considerations, this implies that we need ωNdiff to be much smaller than 1 (order a) and Ndiff to be of the order of 1 implying that the "total attachment/detachment rate" ΩNdiff and the "total diffusion rate" Ndiff should be of the order of the hopping constant ν = 1. i )(cid:105) − (cid:104)nT i (1 − nT i )(cid:105) = ρS i − ρT i .
1206.2744
1
1206
2012-06-13T08:26:46
The capacitance and electromechanical coupling of lipid membranes close to transitions. The effect of electrostriction
[ "physics.bio-ph" ]
Biomembranes are thin capacitors with the unique feature of displaying phase transitions in a physiologically relevant regime. We investigate the voltage and lateral pressure dependence of their capacitance close to their chain melting transition. Since the gel and the fluid membrane have different area and thickness, the capacitance of the two membrane phases is different. In the presence of external fields, charges exert forces that can influence the state of the membrane, thereby influencing the transition temperature. This phenomenon is called electrostriction. We show that this effect allows us to introduce a capacitive susceptibility that assumes a maximum in the melting transition with an associated excess charge. As a consequence, there exist voltage regimes where a small change in voltage can lead to a large uptake of charge and a large capacitive current. Furthermore, we consider electromechanical behavior such as pressure-induced changes in capacitance, and the application of such concepts in biology.
physics.bio-ph
physics
The capacitance and electromechanical coupling of lipid membranes close to transitions. The effect of electrostriction. Thomas Heimburg∗ 1Niels Bohr Institute, University of Copenhagen, Blegdamsvej 17, 2100 Copenhagen Ø, Denmark ABSTRACT Biomembranes are thin capacitors with the unique feature of displaying phase transitions in a physiologically relevant regime. We investigate the voltage and lateral pressure dependence of their capacitance close to their chain melting transition. Since the gel and the fluid membrane have different area and thickness, the capacitance of the two membrane phases is different. In the presence of external fields, charges exert forces that can influence the state of the membrane, thereby influencing the transition temperature. This phenomenon is called electrostriction. We show that this effect allows us to introduce a capacitive susceptibility that assumes a maximum in the melting transition with an associated excess charge. As a consequence, there exist voltage regimes where a small change in voltage can lead to a large uptake of charge and a large capacitive current. Furthermore, we consider electromechanical behavior such as pressure-induced changes in capacitance, and the application of such concepts in biology. ∗corresponding author, [email protected] Keywords: capacitance; voltage; biomembrane; phase transition; piezoelec- tricity, flexoelectricity Abbreviations: DMPC - 1,2 dimyristoyl-3-sn-phosphatidylcholine; DMPG - 1,2 dimyristoyl-3-sn-phosphatidylglycerol; AITC - allylisothiocyanate ; pdf - probability distribution function; HEK cell - human embryonic kidney cell; TRP channel - transient receptor potential channel; ADPR - adenosine diphos- phate ribose drugs or DNA into cells, voltages can be on the order of several 100 mV (5, 6). Stimulation voltages for nerve pulses can be of order up to 1 V (e.g., (7)). A change in voltage leads to a capacitive current because the charge on the capacitor changes. The capacitive current induced by a change in voltage is given as dVm dt + Vm dCm dt (3) dq dt = d dt (Cm · Vm) = Cm In electrophysiological models such as the Hodgkin-Huxley model for nerve pulse propagation and in the interpretation of voltage clamp experiments, it is assumed that the capacitance of biomembranes (in particular of nerves) is independent of volt- age (i.e., Cm is constant) so that the second term on the right of this equation is zero (e.g., (2, 3)). This is equivalent to as- suming that membrane dimensions are unaffected by electri- cal phenomena and that the excitation of membranes does not change their dimensions. The second of these assumptions is known to be incorrect because changes in the thickness of nerve membranes during the action potential have been observed (8 -- 14). Furthermore, there are numerous reports in the literature on voltage induced changes in membrane bending, i.e., caused by flexoelectricity or mechanoelectricity (15 -- 17). In the re- cent years, we have proposed that the voltage changes during the nerve pulse are actually related to changes in capacitance (13, 18 -- 22). In this theoretical paper we show that the assumption of constant capacitance is incorrect, especially if one is close to chain melting transitions in the lipid membrane. Biological membranes display transitions close to physiological temper- ature. Heat capacity maxima are typically found 10-15 degrees below physiological or growth temperature, both for bacterial membranes from E.coli and bacillus subtilis, for lung surfac- tant (13) and for nerves from the spine of rats (S. B. Madsen, N. V. Olsen, unpublished data). The fluid state membrane is thinner than the gel membrane. Simultaneously, the area of the fluid membrane is larger (23). This implies that the capaci- tance of the fluid membrane is larger than that of the gel mem- brane. Thus, any phenomenon in the biomembrane related to a phase transition will influence its capacitance. For instance, the temperature of the phase transition is influenced by charges on the capacitor because electrostatic forces act on the capac- 2 1 0 2 n u J 3 1 ] h p - o i b . s c i s y h p [ 1 v 4 4 7 2 . 6 0 2 1 : v i X r a Introduction Biological membranes provide a barrier between cells and or- ganelles that serves to maintain differences in chemical and electrical potentials by separating molecules and ions. The electrical phenomena which result from transient changes in these electrochemical potentials provide the basis for our present understanding of the electrophysiology of biomembranes (1, 2). The cell membrane consists mainly of a lipid bilayer into which proteins are embedded. It is widely believed that the lipid bi- layer itself is impermeable to water, ions and molecules. There- fore, electrophysiology considers the membrane to be a capaci- tor, and ion channel proteins are regarded as electrical resistors. The nerve pulse, for instance, is considered as a propagating segment of charged capacitor loaded by currents through the channel proteins (3). The capacitance, Cm, defines how much charge, q, is stored on two capacitor plates at a fixed voltage, Vm, q = Cm · Vm . For a parallel plate capacitor, Cm is given by Cm = 0 A D , (1) (2) with vacuum permittivity 0 = 8.85410−12 F/m and dielectric constant  ≈ 2 − 4. A is the area of the membrane, and D is its thickness. Excitatory processes in cells are typically accompanied by changes in voltage. During nerve pulses, for instance, the volt- age changes transiently by about 100 mV in one millisecond. Voltages as large as 100 mV are also typical in the voltage clamp experiments that are used to measure protein conduc- tances (4). In electroporation experiments, used to transport 1 itor plates. Further, hydrostatic pressure (24), lateral pressure (13) or the addition of drugs like anesthetics (25) influence the position of the melting transition. Therefore, the capacitance must be also a function of voltage, hydrostatic pressure, lateral pressure, and the concentration of anesthetics. We show here that close to a transition voltage is able to change the dimensions of a membrane and the capacitance. We also show how the electrical properties of the membrane are affected by the application of lateral pressure or tension in a membrane. This gives rise to pressure-induced capacitive cur- rents or pressure-induced voltage changes. While we restrict ourselves primarily to the phenomenon of electrostriction, i.e., the force that capacitive charges exert on the capacitor, we also discuss polarization effects. Theory Consider a capacitor whose equilibrium properties depend on voltage Vm only, i.e., all other intensive variables of the system such as pressure and temperature are kept constant. We write dq = dVm ≡ CmdVm . (4) (cid:18) ∂q (cid:19) ∂Vm Here we introduce the function Cm = (∂q/∂Vm), which we call the capacitive susceptibility. Note, that the definition of Cm differs from that of the capacitance that is given by Cm = q/Vm. According to eq. (1), eq. (4) corresponds to (cid:18) ∂(CmVm) (cid:19) ∂Vm (cid:18) = CmdVm + VmdCm . dq = dVm = Cm + Vm (cid:19) ∂Cm ∂Vm dVm (5) This equation takes into account that the changes of the charge on a capacitor are not solely due to voltage changes but also to voltage-induced changes in capacitance, which are described by the capacitive susceptibility capacitor plates. In the absence of other forces, increasing the voltage on a capacitor will generally tend to deform the capac- itor such that its thickness is reduced. 1 2 Consider a membrane with fixed thickness D, a capacitance Cm (given by eq. (2)) and a transmembrane voltage, Vm. The field across the membrane is E = Vm/D (assuming a uniform dielectric constant in the membrane interior), and the charge is q = Cm Vm. The force, F, acting on this capacitor is given by (27) Cm V 2 m 1 2 Vm D F = E · q = 1 2 . D q = (7) The force F and the field E are vectors normal to the mem- brane surface. Assuming that the fluid membrane has a capac- itance of Cm = 0.5 µF/cm2 = 0.5 · 10−2 F/m2, this results in a pressure on the membrane of p = 104 N/m2 at 100 mV, which corresponds to 0.1 bar. Due to its quadratic dependence on voltage, this pressure is 100 times larger (p = 10 bar) at Vm = 1 volt. In contrast to hydrostatic pressure this pressure has a direction normal to the membrane. This implies that in- creasing this force results in a reduction of thickness and an increase in area. Since the melting of a membrane is linked to an increase in area, an increase of transmembrane voltage can therefore potentially melt a membrane. At constant voltage, the work done by the electrical field upon melting of the membrane is given by ∆Wc = FdD = 1 2 0V 2 m  A D2 dD , (8) where Dg and Df are the thickness of the gel and the fluid membrane, respectively. For constant area this relation yields m. the familiar expression for the work done on a capacitor, 1 2 CmV 2 However, the membrane area does not stay constant here. The area in the fluid state of DPPC is 24.6% larger than in the gel state and the thickness is 16.3% smaller (23). If we assume a dielectric constant  independent of voltage and mem- brane state, this difference is (cid:90) Df Dg (cid:90) Df Dg Cm ≡ Cm + Vm ∂Cm ∂Vm . (6) C f luid m = 0 Ag · (1 + 0.246) Dg · (1 − 0.163) = 1.49 C gel m . (9) If the capacitance Cm is independent of voltage, we obtain Cm = Cm. However, the last term on the right hand side of eq. (6) can become large close to transitions in biomembranes as we will show below. In the context of transitions, this term can be considered an excess capacitance. It is proportional to the voltage. Therefore it is zero at zero voltage where the ca- pacitor is not charged. In contrast, the function Cm has a finite value since it depends only on the dimensions of the capacitor. The capacitive susceptibility Cm is fully analogous to other susceptibilities such as heat capacity, (dH/dT )p, isothermal volume compressibility, −(dV /dp)T , and isothermal area com- pressibility, −(dA/dΠ)T . The capacitive susceptibility has been used before, e.g., by (26). Electrostriction Electrostriction is the generation of a mechanical force on a ca- pacitor by the electrostatic attraction of the charges on the two Thus, the capacitance in the fluid phase is about 1.5 times larger than that of the gel phase, and the force across the membrane is 1.78 time larger than in the gel phase. In the presence of a voltage difference, a sudden change in membrane state can lead to significant capacitive currents (see discussion). The assumption made above of a voltage- and temperature- independent dielectric constant may not be correct. Paraffin oil has a dielectric constant of 2.2− 4.7 and olive oil has 3.1 while paraffin wax has 2.1 − 2.5 (28). Therefore, it may be possible that the dielectric constant in the fluid membrane is somewhat higher. Lacking reliable data for membranes, we consider  to be constant. Voltage dependence of the melting temperature In the following we describe the melting of a membrane in the presence of a transmembrane voltage. We approximate the excess heat capacity in the absence of voltage, ∆cp,0 = 2 d(∆H0(T ))/dT , by assuming that the transition is governed by a van't Hoff law, i.e., by a two state-transition from gel to fluid with a temperature dependent equilibrium constant K(T ). The temperature dependence of the excess enthalpy, ∆H0(T ) is given by ∆H0(T ) = ∆H0 with K(T ) 1 + K(T ) (cid:18) (cid:18) 1 T (cid:19)(cid:19) K(T ) = exp −n · ∆H0 k − 1 Tm , (10) where n is a cooperative unit size that determines how many lipids undergo a transition at the same time. For DPPC, the total excess enthalpy of the transition is ∆H0 = 39 kJ/mol, and the melting temperature is Tm = 314.2 K. The transition of DPPC unilamellar vesicles is reasonably well described by n = 100. (For details of this calculation see (29).) We de- Figure 1: The influence of electrostriction on the melting of mem- branes. Top: Excess heat capacity of DPPC LUV at three different voltages. Bottom: Change of the transition temperature of DPPC as a function of voltage (see eq. (15)). in both volume and area are proportional to the excess enthalpy during the chain melting transition (23, 24): ∆V (T ) = γV ∆H0(T ) ∆A(T ) = γA∆H0(T ) , (11) where γV = 7.8 · 10−10m3/J is a constant that is practically independent of the lipid species or the lipid mixture (24), and γA = 0.89m2/J. We assume that a similar relation holds for the mean thickness, ∆D(T ) = γD∆H0(T ) , (12) with γD = −2.49 · 10−14m/J. The thickness of the gel phase membrane, Dg, is 4.79 nm (for DPPC), and it is Df = 3.92 nm in the fluid phase. DPPC in the gel phase has an area of Ag = 0.474 nm2 per lipid and a membrane area 1.43· 105 m2 per mol of lipid (23). Here and below we will assume that the above proportionality to the enthalpy is valid. Further, we assume that the voltage dependence of the pure lipid phases is small. According to eq. (8), the enthalpy change, ∆H(T ), at constant voltage, Vm, of the membrane at temperature, T , is given by D(T )2 dD =→  A(T ) ∆H(Vm, T ) = ∆H0(T ) + 1 (cid:82) D(T ) m Dg 2 0 V 2  (Ag+γA∆H0(T )) (Dg+γD∆H0(T ))2 γDd∆H0(T ) . (cid:82) ∆H0(T ) ∆H0(T )+ 1 2 0 V 2 m 0 (13) Making use of (1 + x)−2 ≈ 1 − 2x for small x and assuming constant , we finally obtain (cid:18) ∆H(Vm, T ) = ∆H0(T )· → 1 2 0 γDV 2 m 1+ 1+ 1 2 Ag D2 g (cid:20) (cid:18) γA Ag −2 γD Dg (cid:19) (cid:21)(cid:19) (14) ∆H0(T ) , where ∆H0(T ) is the temperature-dependent enthalpy in the absence of voltage described by eq. (10). One can now deter- mine the temperature dependence of ∆H(Vm, T ). For T (cid:29) Tm, ∆H(Vm, T ) assumes a constant value, which is the ex- It is a quadratic function of voltage: cess heat of melting. m with α0 = −141.7 [J/V2] using ∆H(Vm) = ∆H0 + α0V 2  = 4. ∆H0 is the melting enthalpy of the membrane in the ab- sence of voltage with melting temperature Tm,0 = ∆H0/∆S0, and ∆H(Vm) is the melting enthalpy of the membrane with voltage Vm. The transition temperature Tm in the presence of voltage can be written as Tm = ∆H(Vm) ∆S0 = Tm,0 + = α0 ∆S0 ∆H0 + α0V 2 m m ≡(cid:0)1 + αV 2 ∆S0 V 2 m (cid:1) Tm,0 = (15) note the membrane volume as V (T ) = Vg + ∆V (T ) and the membrane area as A(T ) = Ag + ∆A(T ), where Vg and Ag are the volume and the area of the gel state lipids, and ∆V (T ) and ∆A(T ) are the temperature dependent changes due to melt- ing. In previous publications we have shown that the changes with α = α0/∆H0 = −0.003634 [1/V2], and ∆S0 = ∆H0/Tm,0. The melting profiles and the melting tempera- ture as a function of voltage are shown in Fig.1. One obtains a shift of the transition temperature towards lower temperatures of −11.4 mK for Vm = 100 mV, of −1.14 K for Vm = 1 V and 3 of −114 K for Vm =10 V. The effect of electrostriction on the melting temperature is obviously small at physiological volt- ages. However, it is large for voltages of more than 1 V. The above calculation applies to a symmetric membrane. Ac- cording to the results of (30) the quadratic voltage dependence of Tm can be shifted on the voltage axes when the membrane is asymmetrically charged. In this case, the maximum melting temperature can be achieved for voltages different than zero. Voltage and temperature dependence of the ca- pacitive susceptibility It has been shown that outside of transitions, capacitance changes induced by voltage are small (31 -- 33) but quadratic in voltage. As above we will therefore assume that the voltage dependence of the capacitance in the transition regime is much higher than that of the pure phase. I.e., we assume that Ag and Af as well as Dg and Df display a negligible voltage dependence. The Figure 2: Bottom: Area and thickness changes as a function of voltage at T = 311 K. The temperature, T = 311 K, is below the melting point of DPPC in the absence of voltage. Top: Capaci- tive susceptibility, Cm, of DPPC LUV as a function of voltage. The shaded area indicates the excess charge of the voltage-induced tran- sition. The dashed line is the voltage dependent capacitance, Cm. charge on a membrane at temperature T and voltage Vm is then given by q(Vm, T ) = 0 Agel + ∆A(Vm, T ) Dgel + ∆D(Vm, T ) Vm (16) and at voltage Vm + dV by q(Vm+dVm, T ) = 0 Agel + ∆A(Vm + dVm, T ) Dgel + ∆D(Vm + dVm, T ) ·(Vm+dVm) . (17) where ∆A and ∆D are again proportional to ∆H (cf., eq. (14)). For fixed voltage, the capacitance is a function of tem- perature only. For fixed temperature, it is a function of voltage only. The area and thickness of the membrane at T = 311 K are shown as a function of voltage in Fig. 2 (left). The capacitive susceptibility, Cm, is now given by Cm = dq dVm = q(Vm + dVm, T ) − q(Vm, T ) dVm . (18) If the temperature T is below the melting temperature, Tm,0, voltage can induce a transition in the membrane. Cm displays a maximum at a transition voltage that depends on the experi- mental temperature. This is shown in Fig. 2 (right) for the case of T = 311K. The units are given in absolute units (per mol of lipid) since the area of the membrane is not constant. For com- parison, the specific capacitance in the gel state is 0.74 µF/cm2, and in the fluid state it is 0.93 µF/cm2. The charge on the capacitor is given by q(Vm, T ) = CmdVm . (19) (cid:90) Vm (cid:90) (cid:18) ∂Cm (cid:19) ∂Vm The charge as a function of voltage is shown in Fig. 3 (left) for various temperatures. One can see that the charge undergoes a stepwise change at the transition voltage. We call this change in charge the "excess charge", ∆q0. It is given by (cf., eq. (6)) ∆q0(T ) = Vm Vm dVm , (20) where the integral is from a voltage below to a voltage above the transition. The excess charge corresponds to the shaded peak area in Fig. 2 (right). It has a value of 634 C/mol. One can also determine the capacitive susceptibility, Cm, and the capacitance, Cm, as functions of temperature. This is shown in Fig. 3 (right) for several voltages. For Vm = 0, the capacitive susceptibility and the capacitance are identical. Fluctuations The heat capacity of a membrane at constant pressure is the temperature derivative of the mean enthalpy and is given by d(cid:104)H(cid:105) dT (cid:104)H(cid:105) = with (21) cp = where (cid:104)H(cid:105) is the statistical mean of the enthalpy averaged over all possible microstates of the system with enthalpy Hi. The enthalpies of the microstates are given by: , (cid:80) (cid:80) i Hi exp (−Hi/kT ) i exp (−Hi/kT ) Hi = Ei + pVi + ΠAi + Ψqi + ... (22) 4 Figure 3: Left: Voltage-induced transition of DPPC LUV at four different (constant) temperatures. Top panel: the charge as a functions of voltage. Bottom panel: capacitive susceptibility as a function of voltage. Right: Capacitive susceptibility Cm as a function of temperature for five different voltages. The dashed lines represent the capacitance Cm as a function of temperature. (cid:10)q2(cid:11) − (cid:104)q(cid:105)2 kT similar susceptibility and can be written as Cm = − dq dΨ = dq dVm = ; T, p, Π = const. (26) All of the above susceptibilities are derivatives of exten- sive variables with respect to the conjugated intensive variables. For such susceptibilities, the fluctuations (cid:10)X 2(cid:11) − (cid:104)X(cid:105)2 are quadratic forms and therefore always positive. Heat capacity, volume and area compressibility, and capacitance must always be positive definite functions. For this reason, one also finds that the integrals of the susceptibilities are always be positive: ∆H =(cid:82) T 2 ∆V = −(cid:82) p2 ∆A = −(cid:82) Π2 =(cid:82) Vm,2 ∆q Π1 T1 p1 Vm,1 T dp > 0 cpdT > 0 (cid:104)V (cid:105) κV (cid:104)A(cid:105) κA T dΠ > 0 CmdVm > 0 for for for for T2 > T1 p2 > p1 Π2 > Π1 (27) Vm,2 > Vm,1 . with the intensive variables pressure p, lateral pressure Π, elec- trostatic potential Ψ and the conjugated extensive quantities, internal energy Ei, volume Vi, area Ai, and charge qi. Eq. (21) immediately leads to (cid:10)H 2(cid:11) − (cid:104)H(cid:105)2 kT 2 d(cid:104)H(cid:105) dT = cp = ; p, Π, Ψ, ... = const. (23) which is one of the fluctuation relations. Similar relations are found for the specific isothermal volume compressibility κV T , and the area compressibility κA (cid:18) d(cid:104)V (cid:105) (cid:19) (cid:19) (cid:18) d(cid:104)A(cid:105) dp dΠ T (23): (cid:10)V 2(cid:11) − (cid:104)V (cid:105)2 (cid:10)A2(cid:11) − (cid:104)A(cid:105)2 (cid:104)V (cid:105) kT (cid:104)A(cid:105) kT = = T = − 1 κV (cid:104)V (cid:105) T = − 1 κA (cid:104)A(cid:105) ; T, Π, Ψ = const. (24) ; T, p, Ψ = const. (25) Like the heat capacity, these compressibilities have maxima in the melting transition (23). The capacitive susceptibility of the membrane, Cm, is a This implies that an increase in voltage across a membrane must result in an increase in charge so long as intensive vari- ables other than Vm are kept constant. This is in agreement with the findings in Fig. 3 (left). 5 For derivatives of extensive quantities with respect to non- conjugated quantities, the fluctuations are no longer positive definite forms, and negative values can be obtained. For in- stance, the volume expansion coefficient is given by d(cid:104)V (cid:105) /dT =[(cid:104)V H(cid:105) − (cid:104)V (cid:105)(cid:104)H(cid:105)] /kT 2. For water at 0◦C it is negative. Similarly, derivatives of other extensive variables with respect to non-conjugated intensive variables, such as d(cid:104)q(cid:105) /dT or d(cid:104)q(cid:105) /dΠ, can also be negative (see below). Piezoelectricity A material is said to be "piezoelectric" if the application of a force produces an electric field (and vice vera). "Piezo" orig- inates from the Greek word for pressure, and we will there- fore use the term 'piezoelectric' as synonymous with 'elec- tromechanical' in the sense of pressure-induced voltages across membranes. At fixed temperature, (cid:18) ∂q (cid:19) ∂Vm F (cid:19) (cid:18) ∂q ∂F Vm dq = dVm + dF , (28) where F is the force normal to the membrane. Since thickness changes in the melting transition are coupled to area changes, this leads to the relation (cid:18) ∂q (cid:18) ∂q ∂Vm (cid:19) (cid:19) dq = = dVm + F (cid:18) ∂Π (cid:124) ∂F (cid:19) (cid:123)(cid:122) dΠ Vm Vm dF (cid:125) (cid:18) ∂q (cid:18) ∂q ∂Π (cid:19) (cid:19) dVm + ∂Vm Π ∂Π Vm dΠ , (29) where Π is the lateral pressure of the membrane. In the fol- lowing we will focus on lateral pressure changes. At constant voltage, the enthalpy change, ∆H(T ), of the membrane at tem- perature T due to a lateral pressure is given by a modified ver- sion of eq. (13): ∆H(Vm, T, Π) = ∆H0(T ) + ∆Wc(Vm) + ∆WA(Π) ∆WA(Π) = Π∆A = ΠγA∆H0(T ) , (30) where ∆WA(Π) is the work done to change the area of the membrane from Ag to A. With the help of eq. (14) this leads to · → ∆H(Vm, T, Π) = ∆H0(T )(1 + and the melting temperature is given by (cid:20) 1 − 1 2 (cid:18) γA Agel Agel D2 gel 1 2 + γAΠ) , 0 γDV 2 m (cid:19)(cid:21) m + γAΠ(cid:1) Tm,0 + 2 γD Dgel ∆H0(T ) =(cid:0)1 + αV 2 (31) ∆H(Vm, Π) Tm = (32) with α = −0.003634 [1/V2] (see eq. (15), and γA = 0.89 m2/J. ∆S0 The charge on a membrane at temperature T, voltage Vm and lateral pressure Π can be written as q(Vm, T, Π) = 0 Agel + ∆A(Vm, T, Π) Dgel + ∆D(Vm, T, Π) Vm (33) 6 Figure 4: Changes in electrical properties induced by lateral pres- sure changes. Top: Change in the charge on a capacitor upon changes in lateral pressure at fixed voltage of Vm =1 V. Bottom: Change in voltage induced by lateral pressure changes at fixed charge on the capacitor (Cm=1430.7 C/mol corresponding to the charge on the fluid phase membrane at Vm=1 V). Increasing lateral pressure renders the membrane more solid. where ∆A and ∆D are again proportional to ∆H (cf. eq. (31)) Now we can determine (1) how the charge on the capacitor changes with changes in lateral pressure at constant voltage and temperature (and vice versa) and (2) how the voltage changes at constant temperature and constant charge with changes in lateral pressure (and vice versa). Fig. 4 (left) shows case 1 for a fixed voltage of Vm= 1 V and T = 315 K. Under these conditions, the membrane is in the fluid state when the lateral pressure Π is zero. Increasing pressure renders the membrane more solid and the capacitance of the membrane decreases. This leads to a release of charge from the capacitor (i.e., pressure induced capacitive currents). One can define the corresponding susceptibility (cid:18) ∂q (cid:19) ∂Π Vm βV ≡ , (34) which is shown in Fig. 4 (right). This susceptibility represents the change in the charge on the membrane due to an increment in the pressure at constant voltage. Since it is a derivative of the extensive variable q with respect to the non-conjugated exten- sive variable Π, it can have a negative value. Discussion Lipid membranes are thin capacitors that are unique in their property to display transitions in capacitance. Here, we have offered a theoretical framework for describing the capacitance and the capacitive susceptibility of membranes in the transi- tion regime. Since biomembranes are close to such transitions, this is of immediate biological relevance. Charges on mem- branes create forces on the membrane that tend to compress the membrane normal to the membrane surface. This effect is called electrostriction (27). It leads to a lowering of the melting temperature of the membrane. Therefore, a change of trans- membrane voltage can induces melting transitions. As a conse- quence, there is an excess charge linked to the voltage-induced transition. This is expressed by the capacitive susceptibility, Cm = ∂q/∂Vm, which displays a pronounced maximum at the transition. This behaviour is very similar to that of the heat capacity as a function of temperature and the volume and area compressibilities as a function of hydrostatic and lateral pres- sure. These functions are all derivatives of extensive thermody- namic variables with respect to their conjugated intensive vari- able. Using the fluctuation theorem, we have shown that the capacitive susceptibility is related to fluctuations in charge and must therefore always be positive. The maximum of the capac- itive susceptibility describes the fact that in particular voltage regimes close to transitions a small change in voltage may lead to a large uptake of charge, i.e., the membrane is very suscepti- ble to small variations in conditions. We further show that there exist electromechanical cou- plings in lipid membranes. One finds a straight-forward con- nection between lateral pressure and the charge on a membrane (at constant voltage) that also assumes a maximum in the melt- ing transition of the membrane. This implies that changes in pressure can give rise to capacitive currents. This effect is de- scribed by the derivative of an extensive variable (charge) with respect to a non-conjugated variable (lateral pressure). Such susceptibilities may have either positive and negative values. At constant charge, there also exists a coupling between volt- age and lateral pressure. This effect corresponds to piezoelec- tricity. The coupling constant assumes a maximum in the melt- ing transition. This effect is very important for excitatory pro- cesses (for instance nerve pulse) and is discussed below. In the present paper, we have not considered other potentially signifi- cant effects such as field-induced changes in polarization. Such effects are relevant and clearly exist in lipid monolayers. This will be the focus of a future publication. Due to their different capacitance, gel and fluid membranes carry different charges when a constant voltage is applied. An interesting consequence is that the presence of a transmem- brane voltage will lead to charge separation in the melting regi- me where gel and fluid domains coexists. If the membranes contain a fraction of charged lipids, this will lead to an asym- metric accumulation of charged lipids in the fluid domains. Thus, the shape of phase diagrams will be influenced by the application of an electrical field across the membrane. Not sur- prisingly, besides temperature, pressure and the concentrations of the components, the transmembrane voltage is also a vari- able that determines phase diagrams and should be included in Gibbs' phase rule. The influence of voltage on the capacitance of black lipid Figure 5: Top: Change in voltage induced by lateral pressure changes at fixed charge on the capacitor (Cm=1430.7 C/mol cor- responding to the charge on the fluid phase membrane at Vm=1 V). Increasing lateral pressure renders the membrane more solid. Bot- tom: The corresponding susceptibility, βq = (∂Vm/∂Π)q. When the charge in eq. (28) is kept constant (dq = 0), we obtain − (cid:16) ∂q (cid:17) (cid:17) (cid:16) ∂q ∂Π Vm  dΠ ≡ βqdΠ . ∂Vm Π dVm = (35) (36) This is shown in Fig. 5 (left) where the charge is fixed at q =1645.4 C/mol, which is the charge on the fluid membrane at a voltage of Vm = 1 V. The work done on the membrane by pressure is converted into work done on the charges by chang- ing their distance. Increasing pressure at fixed charge leads to a larger voltage across the membrane. We call this a piezoelectric effect. The coupling constant, (cid:18) ∂Vm (cid:19) ∂Π q βq = is shown in Fig. 5 (right). It is identical to the term in rectan- gular brackets in eq. (35). A situation of constant charge could be present in a membrane during a sudden reversible change in lateral pressure when charges have no time to dissociate. 7 membranes had been investigated both theoretically and ex- perimentally in several studies in the 1970's (30, 31, 34 -- 36). All of these studies assume that electrostriction is the domi- nant effect leading to a reduction in membrane thickness and an increase in area due to an increase in voltage. Those studies were done far away from the phase transitions temperature of black lipid membranes. In agreement with our derivations here, they found that the dependence of the capacitance should be a quadratic function of voltage, i.e., Cm ∝ (1 + αV 2 m), where α is a constant. This constant is strongly influenced by the pres- ence of solvent in the black lipid membranes (32). Solvent-free membranes display a much smaller voltage dependence of the capacitance on voltage. In an interesting experimental study, Alvarez and Latorre (30) found that changes in capacitance in asymmetric membranes are influenced by a resting potential, E0, such that Cm ∝ (1 + α(V − E0)2). Alvarez and Latorre discussed this finding in the context of nerve pulse propagation and the measurement of gating currents. They suggested that the membrane itself could display capacitive currents similar to gating currents. The novel aspect of the present paper is the recognition of the profound effect that the melting transition has on the nonlinear behavior of the membrane capacitance. While previous authors have assumed a constant compressibil- ity of the membrane, we have made use of the fact that the compressibility is dramatically increased in transitions. The two terms of the capacitive current Cm · dVm/dt and Vm · dCm/dt in eq. (3) can have different time dependences. The first term is fast and is largely determined by the electri- cal resistance of the aqueous medium. The second term has a time scale given by the relaxation time of the membrane, which is slow in transitions. It should therefore be possible to distin- guish the two contributions to the capacitive current experimen- tally. We have previously shown that relaxation time scales can be as large as 30 seconds for artificial membranes at the tran- sition maximum, and we estimate time scales of up to 100ms for biomembranes (37, 38). Besch and collaborators investi- gated HEK293 cells and found currents with timescales in the 10ms regime after voltage change (39), which indicates that these currents are related to time-dependent changes in mem- brane geometry rather than being caused by charging a mem- brane with a constant capacitance. Slow capacitive currents after voltage changes were also found in rat nerves (40) but in- terpreted as gating currents. The electromechanical aspect of membranes has received considerable interest in the past. Ochs and Burton showed in 1974 that black lipid membranes made of egg-PC and choles- terol show electromechanical behavior under voltage clamp con- ditions (15). They applied an oscillating pressure difference across the membrane and recorded capacitive currents that they described by IC = Vm · dCm/dt (cf., eq. (3)). As mentioned above, the interesting issue of polarization has not been treated here. However, much of the literature on electromechanical phenomena in membranes is dedicated to polarization caused by membrane curvature. R. B. Meyer proposed in 1969 that liquid crystals of molecules that possess electrical dipole moments should also be piezoelectric (41). In particular he stated that "another possible application of these effects may be in interpreting some of the potentials and ion distributions observed in liquid-crystalline biological structures." Following this idea, Petrov and collaborators proposed that mem- 8 branes should also display piezoelectric behavior (42). Lipid monolayers have significant dipole moments due to the polar head groups and the oriented associated water layer. For a sym- metric membrane of a zwitterionic lipid, the polarizations of the two layers should cancel because they have opposite ori- entations. However, in the presence of curvature, this sym- metry argument does not apply, and one expects an electrical field generated by curved membranes. Petrov named this phe- nomenon "flexoelectricity" and demonstrated the effect in ex- periments similar to those by Ochs and Burton (15) but inter- preted differently as curvature-induced polarization (43, 44). In several further papers the authors applied the concept of flexoelectricity to biomembranes and proposed a coupling of flexoelectricity to ionic currents through channel proteins (45, 46). Since lipid membranes in the absence of proteins also dis- play ion-channel-like characteristics close to transitions (47 -- 49), Petrov's considerations are also valid here. It is reason- able to expect that the polarization, P , of membranes changes significantly in the phase transition regime and that the related susceptibility, dP/dE (with E being the electrical field), should have a maximum in the transition. This must be so since the dipole moment of a gel monolayer differs from that of a fluid monolayer. Upon membrane bending one should find an asym- metric distribution of gel and fluid lipids in the two oppos- ing monolayers (50) leading to an effective polarization of the membrane. Unfortunately, we are not aware of any reliable ex- perimental data or theoretical estimates of the magnitude of this polarization. To our knowledge, the above authors did not in- vestigate the effect of the phase transition on flexoelectricity. However, polarizations induced by curvature are completely analogous to voltage changes induced by lateral compression as described above by eq. Interestingly, Helfrich investigated the effect of voltage on the phase tran- sition temperature of 3-dimensional liquid crystals as early as 1970 (51). He found that the shift of the transition is given by 2 Tm,00∆E2/∆H0ρ, where ∆ is the difference of ∆Tm = 1 the dielectric constant between liquid and solid phase, ρ is is the density and E is the electric field. Assuming E/D = Vm, this law is analogous to eq. 15. (35) and Fig. 5. It is known experimentally that lipid monolayers have large dipole potentials of order 300mV -- somewhat higher for gel than for fluid phase monofilms (52, 53). This is frequently at- tributed to the dipolar nature of the lipid head group and the associated water. One can influence the state of lipid monolay- ers in experiments by an applied field. At positive voltages a liquid-expanded monolayer becomes more solid, while the op- posite effect is observed when the field is reversed. This rules out the possibility that the influence of voltage is due to elec- trostriction (PhD Kasper Feld, NBI 2012), which is indepen- dent of the direction of the field. For membranes, this is less clear. Antonov and collaborators (54) measured the voltage dependence of the lipid chain melting transition via the effect of voltage on membrane permeability, which displays a maxi- mum at the melting point. For synthetic black lipid membranes made of either DPPC or DPPA (dipalmitoyl phosphatidic acid), they found an increase in melting temperature that was well de- scribed by Tm(DPPC) = 315.4[K] + 20.5[K/V ] · Vm , Tm(DPPA) = 332.8[K] + 55.1[K/V ] · Vm . (37) This corresponds to a linear shift of Tm of +1.03 K at Vm = 50mV for DPPC and a shift of 2.75 K for DPPA towards higher temperatures, respectively. This effect is opposed to the trend towards lower temperatures predicted above based on electrostric- tion, which is in agreement with previous predictions from Sugar (55). However, Antonov's finding is remarkably close to the calculation of Cotterill (56) made by considering the polariza- tion of the monolayers. The shift of Tm by Antonov was found to be linear within experimental accuracy as expected from Cot- terill's calculation. We do not believe, however, that the deriva- tion of Cotterill is theoretically sound. An increase of the transition temperature with increasing volt- age poses an interesting theoretical problem. Under such cir- cumstances, the fluid membrane can be made solid by voltage leading to a lower capacitance. This would render the excess capacitive susceptibility in the transition negative an would re- sult in a negative excess charge. According to the considera- tions of fluctuations above, this can hold only if there are cou- plings between the charge on the membrane and non-conjugated intensive variables. It remains to be seen whether the simulta- neous change of capacitance and polarization allows for such behavior. However, if this were the case, an increase in voltage could possibly result in a decrease of charge, i.e., in capacitive currents against the applied field. It should also be noted that the publication of Antonov and collaborators (54) is the only one on voltage-induced shifts in transition temperature that we are aware of, and independent experiments verification would be useful. In this respect, the shift of the quadratic dependence on voltage due to a pre-existing polarization found by Alvarez and Latorre (30) might contain the answer to the present prob- lem. This suggests that the membranes in the experiments of Antonov were not symmetric. Additional studies of polarized membranes are to be encouraged. (9). The biological importance of electromechanical coupling was discussed in connection with the function of the hair cells of the outer ear (57, 58). In a recent paper, Brownell and col- laborators showed that tethers pulled from hair cells contract upon application of a voltage difference across the cell body (59). The phenomena discussed in the present publication are especially important for nerve pulse propagation. A change of the membrane state from fluid to gel will alter the capacitance by about 50%,see eq. If this change occurs within 0.5 ms (which is the time scale of the rising phase of the nervous impulse), it will generate a significant capacitive current of or- der 100 µA/cm2 at a constant voltage of Vm = 100 mV. Ionic currents of a similar order of magnitude are central elements in the Hodgkin-Huxley model of the nerve pulse (3) (Fig. 18 therein). Hodgkin and Huxley calculated net ion currents of the order of 100-600 µA/cm2 in the squid axon. In previous pub- lications we have suggested that the nervous impulse consists of an electromechanical soliton corresponding to a lateral com- pression of the neuronal membranes (13, 18 -- 22). In particular, we have proposed that the membrane undergoes a change in state from fluid to gel and back during the nerve pulse. Thus, in the soliton theory one expects capacitive currents of similar magnitude as the ionic currents in the Hodgkin-Huxley model. If charges cannot dissociate (i.e., no capacitive current), one rather expects changes in transmembrane potential (Fig. 5). The consequence of these non-linear effects is an electrome- chanical soliton that can travel along membrane cylinder with 9 a velocity close to the speed of sound with many similarities to the nervous impulse. In fact, Tasaki and collaborators showed in various publications that such mechanical pulses exist (8 -- 12, 60). Further support comes from atomic force experiments, that show mechanical signals in synapse in phase and propor- tional to voltage changes (14). Further, light scattering tech- niques demonstrate that voltage changes are accompanied by changes in nerve dimensions (61, 62). One might reasonably assume that polarization pulses can also propagate in membranes. In lipid monolayers, such pres- sure pulses have recently been demonstrated experimentally by Schneider and collaborators (63). The pressure pulses are ac- companied by voltage pulses that are directly related to and in phase with the pressure pulse (M. F. Schneider, private com- munication). The coupling displays a maximum in the phase transition regime in an agreement with the concepts proposed here. Such experiments are important and will eventually lead to a complete thermodynamic picture of the capacitive suscep- tibility and the electromechanical behavior of biomembranes and nerves. Conclusion We have shown here that lipid membranes are very susceptible to voltage changes close to phase transitions. This result ex- tends previous theoretical and experimental findings of voltage- dependent capacitances of artificial membranes far from such transitions. We have introduced a capacitive susceptibility that displays a pronounced maximum at the transition. Voltage chan- ges generate changes in area that result in an electromechanical coupling. Since biomembranes are exist naturally in a state close to a transition, this effect will play a role in excitatory processes of the cell. One important example is nerve pulse propagation. Acknowledgments: I acknowledge communication with Dr. K. Vanselow from Kiel, who explored the influence of voltage on the mechanics of membranes as early as 1963. I thank Prof. A. D. Jackson from the NBI for critical reading and helpful comments. The support of the Villum Foundation (Denmark) is gratefully acknowledged. References 1. Hille, B., 1992. Ionic channels of excitable membranes. Cambridge Uni- 2. Johnston, D., and S. M. S. Wu, 1995. Cellular Neurophysiology. MIT versity Press, Cambridge. Press, Boston. 3. Hodgkin, A. L., and A. F. Huxley. 1952. A quantitative description of membrane current and its application to conduction and excitation in nerve. J. Physiol. London 117:500 -- 544. 4. Hodgkin, A. L., and A. F. Huxley. 1952. Currents carried by sodium and potassium ions through the membrane of the giant axon of loligo. J. Physiol. London 116:449 -- 472. 5. Neumann, E., S. Kakorin, and K. Taensing. 1999. Fundamentals of electro- porative delivery of drugs and genes. Bioelectrochem. Bioenerg. 48:3 -- 16. 6. Gehl, J. 2003. Electroporation: theory and methods, perspectives for drug delivery, gene therapy and research. Acta Physiol. Scand. 177:437 -- 447. 7. Kassahun, B. T., A. K. Murashov, and M. Bier. 2010. A thermodynamic mechanism behind an action potential and behind anesthesia. Biophys. Rev. Lett. 5:35 -- 41. 8. Iwasa, K., and I. Tasaki. 1980. Mechanical changes in squid giant-axons lipid membranes and its relation to the heat capacity. Biophys. J. 82:299 -- 309. 38. Seeger, H. M., M. L. Gudmundsson, and T. Heimburg. 2007. How anesthetics, neurotransmitters, and antibiotics influence the relaxation pro- cesses in lipid membranes. J. Phys. Chem. B 111:13858 -- 13866. 39. Besch, S., K. V. Snyder, R. C. Zhang, and F. Sachs. 2003. Adapting the quesant nomad atomic force microscope for biology and patch-clamp atomic force microscopy. Cell Biochem. Biophys. 39:195 -- 210. 40. Kilic, G., and M. Lindau. 2001. Voltage-dependent membrane capacitance in rat pituitary nerve terminals due to gating currents. Biophys. J. 80:1220 -- 1229. 41. Meyer, R. B. 1969. Piezoelectric effects in liquid crystals. Phys. Rev. Lett. 22:918 -- 921. 42. Petrov, A. G. 1984. Flexoelectricity of lyotropics and biomembranes. Nuovo Cimento D 3:174 -- 192. 43. Petrov, A. G., and V. S. Solokov. 1989. Curvature-electric effects in arti- ficial and natural membranes studied using patch-clamp techniques. Eur. Biophys. J. 17:139 -- 155. 44. Petrov, A. G., and P. N. R. Usherwood. 1994. Mechanosensitivity of cell membranes. Eur. Biophys. J. 23:1 -- 19. 45. Petrov, A. G., B. A. Miller, K. Hristova, and P. N. R. Usherwood. 1993. Flexoelectric effects in model and native membranes containing ion chan- nels. Eur. Biophys. J. 22:289 -- 300. 46. Petrov, A. G. 1997. Charge transfer processes in model and biological membranes: Defect and mechano-electricaspects; statics and dynamics. Mol. Cryst. Liq. Cryst. A 292:227 -- 234. 47. Blicher, A., K. Wodzinska, M. Fidorra, M. Winterhalter, and T. Heimburg. 2009. The temperature dependence of lipid membrane permeability, its quantized nature, and the influence of anesthetics. Biophys. J. 96:4581 -- 4591. 48. Heimburg, T. 2010. Lipid ion channels. Biophys. Chem. 150:2 -- 22. 49. Laub, K. R., K. Witschas, A. Blicher, S. B. Madsen, A. Luckhoff, and T. Heimburg. 2012. Comparing ion conductance recordings of synthetic lipid bilayers with cell membranes containing trp channels. Biochim. Bio- phys. Acta 1818:1 -- 12. 50. Ivanova, V. P., and T. Heimburg. 2001. A histogram method to obtain heat capacities in lipid monolayers, curved bilayers and membranes containing peptides. Phys. Rev. E 63:1914 -- 1925. 51. Helfrich, W. 1970. Effect of electric fields on temperature of phase tran- sitions of liquid crystals. Phys. Rev. Lett. 24:201 -- 203. 52. Vogel, V., and D. Mobius. 1988. Local surface potentials and electric dipole moments of lipid monolayers: contributions of the water/lipid and the lipid/air interfaces. J. Coll. Interf. Sci. 126:408 -- 420. 53. Lee, K. Y. C., and H. M. McConnell. 1995. Effect of electric-field gradi- ents on lipid monolayer membranes. Biophys. J. 68:1740 -- 1751. 54. Antonov, V. F., E. Y. Smirnova, and E. V. Shevchenko. 1990. Electric field increases the phase transition temperature in the bilayer membrane of phosphatidic acid. Chem. Phys. Lipids 52:251 -- 257. 55. Sugar, I. P. 1979. A theory of the electric field-induced phase transition of 56. Cotterill, R. M. J. 1978. Field effect on lipid melting. Phys. Scripta 18:191 -- 192. 57. Rabbitt, R. D., H. E. Ayliffe, D. Christensen, K. Pamarth, C. Durney, S. Clifford, and W. E. Brownell. 2005. Evidence of piezoelectric reso- nance in isolated outer hair cells. Biophys. J. 88:2257 -- 2265. 58. Sachs, F., W. E. Brownell, and A. G. Petrov. 2009. Membrane electrome- chanics in biology, with a focus on hearing. MRS Bull. 34:665 -- 670. 59. Brownell, W. E., F. Qian, and B. Anvari. 2010. Cell membrane tethers generate mechanical force in response to electrical stimulation. Biophys. J. 99:845 -- 852. 60. Tasaki, I., and M. Byrne. 1990. Volume expansion of nonmyelinated nerve fibers during impulse conduction. Biophys. J. 57:633 -- 635. 61. Tasaki, I., A. Watanabe, R. Sandlin, and L. Carnay. 1968. Changes in fluorescence, turbidity and birefringence associated with nerve excitation. Proc. Natl. Acad. Sci. USA 61:883 -- 888. 62. Rector, D. M., X. Yao, R. M. Harper, and J. S. George, 2009. In vivo observations of rapid scattered light changes associated with neurophys- iological activity. In In vivo optical imaging of brain function (Frostig, R. D., editor). CRC Press, Boca Raton, FL, 143 -- 170. 63. Griesbauer, J., S. Bossinger, A. Wixforth, and M. F. Schneider. 2012. Propagation of 2d pressure pulses in lipid monolayers and its possible im- plications for biology. Phys. Rev. Lett. 108:198103. 26. Carius, W. 1976. Voltage dependence of bilayer membrane capacitance. phospholipid bilayers. Biochim. Biophys. Acta 556:72 -- 85. associated with production of action potentials. Biochem. Biophys. Re- search Comm. 95:1328 -- 1331. 9. Iwasa, K., I. Tasaki, and R. C. Gibbons. 1980. Swelling of nerve fibres associated with action potentials. Science 210:338 -- 339. 10. Tasaki, I., K. Iwasa, and R. C. Gibbons. 1980. Mechanical changes in crab nerve fibers during action potentials. Jap. J. Physiol. 30:897 -- 905. 11. Tasaki, I., and K. Iwasa. 1982. Further studies of rapid mechanical changes in squid giant axon associated with action potential production. Jap. J. Physiol. 32:505 -- 518. 12. Tasaki, I., K. Kusano, and M. Byrne. 1989. Rapid mechanical and ther- mal changes in the garfish olfactory nerve associated with a propagated impulse. Biophys. J. 55:1033 -- 1040. 13. Heimburg, T., and A. D. Jackson. 2005. On soliton propagation in biomembranes and nerves. Proc. Natl. Acad. Sci. USA 102:9790 -- 9795. 14. Kim, G. H., P. Kosterin, A. Obaid, and B. M. Salzberg. 2007. A mechani- cal spike accompanies the action potential in mammalian nerve terminals. Biophys. J. 92:3122 -- 3129. 15. Ochs, A. L., and R. M. Burton. 1974. Electrical response to vibration of a lipid bilayer membrane. Biophys. J. 14:473 -- 489. 16. Raphael, R. M., A. S. Popel, and W. E. Brownell. 2000. A membrane bending model of outer hair cell electromotility. Biophys. J. 78:2844 -- 2862. 17. Petrov, A. G. 2006. Electricity and mechanics of biomembrane systems: flexoelectricity in living membranes. Anal. Chim. Acta 568:70 -- 83. 18. Heimburg, T., and A. D. Jackson. 2007. On the action potential as a propagating density pulse and the role of anesthetics. Biophys. Rev. Lett. 2:57 -- 78. 19. Heimburg, T., and A. D. Jackson, 2008. Thermodynamics of the nervous impulse. In Structure and Dynamics of Membranous Interfaces. (Nag, K., editor). Wiley, 317 -- 339. 20. Andersen, S. S. L., A. D. Jackson, and T. Heimburg. 2009. Towards a thermodynamic theory of nerve pulse propagation. Progr. Neurobiol. 88:104 -- 113. 21. Villagran Vargas, E., A. Ludu, R. Hustert, P. Gumrich, A. D. Jackson, and T. Heimburg. 2011. Periodic solutions and refractory periods in the soliton theory for nerves and the locust femoral nerve. Biophys. Chem. 153:159 -- 167. 22. Lautrup, B., R. Appali, A. D. Jackson, and T. Heimburg. 2011. The stability of solitons in biomembranes and nerves. Eur. Phys. J. E 34:57. 23. Heimburg, T. 1998. Mechanical aspects of membrane thermodynam- ics. Estimation of the mechanical properties of lipid membranes close to the chain melting transition from calorimetry. Biochim. Biophys. Acta 1415:147 -- 162. 24. Ebel, H., P. Grabitz, and T. Heimburg. 2001. Enthalpy and volume changes in lipid membranes. i. the proportionality of heat and volume changes in the lipid melting transition and its implication for the elastic constants. J. Phys. Chem. B 105:7353 -- 7360. 25. Heimburg, T., and A. D. Jackson. 2007. The thermodynamics of general anesthesia. Biophys. J. 92:3159 -- 3165. J. Coll. Interf. Sci. 57:301 -- 307. 27. Vanselow, K. Untersuchungen uber die elektrostatisch- mechanischen kraftean der nervenmembran bei der auslosung des ak- tionspotentials. Z. Dt. Gesellsch. med. biol. Elektronik 11:1 -- 5. 1966. 28. ASI Instruments, 2012. Dielectric Constants Chart. Web site (http://www.asiinstr.com/technical/Dielectric%20Constants.htm). 29. Heimburg, T., 2007. Thermal biophysics of membranes. Wiley VCH, Berlin, Germany. 30. Alvarez, O., and R. Latorre. 1978. Voltage-dependent capacitance in lipid bilayers made from monolayers. Biophys. J. 21:1 -- 17. 31. White, S. H. 1974. Comments on "Electrical breakdown of bimolecular lipid membranes as an electromechanical instability". Biophys. J. 14:155 -- 158. 32. Requena, J., D. A. Haydon, and S. B. Hladky. 1975. Lenses and the compression of black lipid membranes by an electric field. Biophys. J. 15:77 -- 81. 33. Farrell, B., C. Do Shope, and W. E. Brownell. 2006. Voltage-dependent capacitance of human embryonic kidney cells. Phys. Rev. E 73:041930 -- 1 -- 041930 -- 17. 34. White, S. H. 1970. A study on lipid bilayer membrane stability using precise measurements of specific capacitance. Biophys. J. 10:1127 -- 1148. 35. White, S. H., and T. E. Thompson. 1973. Capacitance, area, and thickness variations in thin lipid films. Biochim. Biophys. Acta 323:7 -- 22. 36. White, S. H., and W. Chang. 1981. Voltage dependence of the capacitance and area of blacklipid membranes. Biophys. J. 36:449 -- 453. 37. Grabitz, P., V. P. Ivanova, and T. Heimburg. 2002. Relaxation kinetics of 10
1307.3045
1
1307
2013-07-11T10:34:42
Lipid ion channels and the role of proteins
[ "physics.bio-ph", "cond-mat.soft", "q-bio.BM" ]
Synthetic lipid membranes in the absence of proteins can display quantized conduction events for ions that are virtually indistinguishable from those of protein channel. By indistinguishable we mean that one cannot decide based on the current trace alone whether conductance events originate from a membrane, which does or does not contain channel proteins. Additional evidence is required to distinguish between the two cases, and it is not always certain that such evidence can be provided. The phenomenological similarities are striking and span a wide range of phenomena: The typical conductances are of equal order and both lifetime distributions and current histograms are similar. One finds conduction bursts, flickering, and multistep-conductance. Lipid channels can be gated by voltage, and can be blocked by drugs. They respond to changes in lateral membrane tension and temperature. Thus, they behave like voltage-gated, temperature-gated and mechano-sensitive protein channels, or like receptors. Lipid channels are remarkably under-appreciated. However, the similarity between lipid and protein channels poses an eminent problem for the interpretation of protein channel data. For instance, the Hodgkin-Huxley theory for nerve pulse conduction requires a selective mechanism for the conduction of sodium and potassium ions. To this end, the lipid membrane must act both as a capacitor and as an insulator. Non-selective ion conductance by mechanisms other than the gated protein-channels challenges the proposed mechanism for pulse propagation. ... Some important questions arise: Are lipid and protein channels similar due a common mechanism, or are these similarities fortuitous? Is it possible that both phenomena are different aspects of the same phenomenon? Are lipid and protein channels different at all? ... (abbreviated)
physics.bio-ph
physics
Lipid ion channels and the role of proteins Lars D. Mosgaard and Thomas Heimburg∗ 1Niels Bohr Institute, University of Copenhagen, Blegdamsvej 17, 2100 Copenhagen Ø, Denmark ABSTRACT Synthetic lipid membranes in the absence of proteins can display quantized conduction events for ions that are virtually indistinguishable from those of protein channel. By indistinguishable we mean that one cannot decide based on the current trace alone whether conductance events originate from a membrane which does or does not contain channel proteins. Additional evidence is required to distinguish between the two cases, and it is not always certain that such evidence can be provided. The phenomenological similarities are striking and span a wide range of phenomena: The typical conductances are of equal order and both lifetime distributions and current histograms are similar. One finds conduction bursts, flickering, and multistep-conductance. Lipid channels can be gated by voltage, and can be blocked by drugs. They respond to changes in lateral membrane tension and temperature. Thus, they behave like voltage-gated, temperature-gated and mechano-sensitive protein channels, or like receptors. Lipid channels are remarkably under-appreciated. However, the similarity between lipid and protein channels poses an eminent problem for the interpretation of protein channel data. For instance, the Hodgkin-Huxley theory for nerve pulse conduction requires a selective mechanism for the conduction of sodium and potassium ions. To this end, the lipid membrane must act both as a capacitor and as an insulator. Non-selective ion conductance by mechanisms other than the gated protein-channels challenges the proposed mechanism for pulse propagation. Nevertheless, the properties of the lipid membrane surrounding the proteins hardly ever enter into the textbook discussion of membrane models. Some important questions arise: Are lipid and protein channels similar due a common mechanism, or are these similarities fortuitous? Is it possible that both phenomena are different aspects of the same phenomenon? Are lipid and protein channels different at all? In this review we will document some of the experimental and theoretical findings that show the similarity between lipid and protein channels. We discuss important cases where protein channel function is strongly correlated to lipid properties. Based on some statistical thermodynamics simulations we discuss how such a correlations could come about. We suggest that proteins can in principle act as catalysts for lipid channel formation and that some apparently mysterious correlations between protein and lipid membrane function can be understood in this manner. ∗corresponding author, [email protected] Keywords: membranes, heat capacity, adiabatic compressibility, frequency de- pendence, relaxation, dispersion periments on synthetic membrane suggest that the spontaneous appearance of pores in membranes is related to thermal fluctu- ations, which are known to approach a maximum in the tran- sition regime. The likelihood of pore opening and the respec- tive open lifetimes are characterized well by the well-known fluctuation-dissipation theorem (FDT)(5) and its applications to membrane pores (1). In essence, the FDT establishes con- nections between the fluctuations of extensive variables such as enthalpy, volume and area, and the conjugated susceptibilities: heat capacity, isothermal volume and area compressibility, re- spectively. For instance, wherever the heat capacity is high, the fluctuations in enthalpy are large. Similarly, large volume and area fluctuations imply a high volume or lateral compressibil- ity. Volume and area fluctuations are strongly coupled to the enthalpy fluctuations (4). Therefore, the membrane becomes highly compressibly (i.e., soft) close to transition, and the for- mation of defects is facilitated. For this reason, close to tran- sitions one expects the spontaneous formation of pores in the lipid membrane, and a significant increase of the permeabil- ity for ions and small molecules. Such changes in permeabil- ity close to transitions were in fact observed experimentally(1). The presence of melting transitions in biological membranes therefore makes it highly likely that lipid ion channels are also present in living cells. An important implication of the FDT is that fluctuations are directly coupled to fluctuation lifetimes (6), which implies that the open lifetime of lipid membrane pores is maximum close to the transition. Changes in experimental conditions can shift the melting point. For instance, drugs such as the insecticide lindane or the anesthetic octanol lower transition temperatures and therefore influence the permeability of membranes (1). Due to their ef- Introduction 1 The observation of channel-like conduction events in pure lipid membranes is not new but has not received appropriate atten- tion. Yafuso et al. described them in oxidized cholesterol membranes as early as 1974. Further evidence was provided by Antonov and collaborators and by Kaufmann and Silman in the 1980s. It was shown that channel formation is influ- enced by temperature and pH. Goegelein and Koepsell showed in 1984 that one can block such lipid channels with calcium, and Blicher et al. showed that one can block lipid membrane channels with general anesthetics. They may be mildly selec- tive for ions following the Hofmeister series (the above is re- viewed in Heimburg, 2010 (1)). Further, lipid channels can be gated by voltage (2). In the last decade, Colombini and collab- orators described channel-like events in membranes containing ceramides and sphingolipids (3). Synthetic lipid membranes can display melting transitions of their chains (4). The transition is accompanied by an ab- sorption of heat, and a change in entropy due to the disorder- ing of the lipid chains. Biological membranes possess such melting transitions, too, typically at temperatures about 10 de- grees below physiological temperature (4). Chain melting has been described for several bacterial membranes but also for lung surfactant and nerve membranes. Electrophysiological ex- 1 3 1 0 2 l u J 1 1 ] h p - o i b . s c i s y h p [ 1 v 5 4 0 3 . 7 0 3 1 : v i X r a 2 Lipid ion channels 2.1 Single channels in protein-free membranes Fig. 1A shows a typical channel event in a synthetic lipid mem- brane (DMPC: DLPC=10:1, 150mM NaCl) recorded around 30◦C (10). These channels display a conductance of about 330 pS and lifetimes on the order of 1-100ms. Such conductances and lifetimes are not untypical for protein channels, too. In fact, it would be difficult to distinguish the traces in Fig. 1 from protein channels in the absence of independent informa- tion available. Further examples were reviewed in Heimburg, 2010 (1). Fig. 1D shows recordings from so-called ceramide chan- nels. Such channels have been investigated in much detail by the Colombini group from the University of Maryland. Channel events in protein-free membranes containing a small amount of ceramide lipids look similar to the lipid pores in Fig. 1. Ce- ramide channels seem to be distinct from other lipid membrane pores since they display much larger conductances and much longer lifetimes than the data in Fig. 1 (A-C). 2.2 Comparison of lipid and TRP protein ion channels The literature contains copious examples for protein channel conductance. As mentioned, many of these data are similar in appearance to the lipid channels. This is demonstrated in the following for TRP channels that were over-expressed in hu- man embryonic kidney (HEK) cells (10). Fig. 2 (top) shows short but representative time-segments of channel recordings from a synthetic lipid preparation (top trace), HEK cells con- taining the TRPM2 (center trace) and TRPM8 (bottom trace) channels (data from Laub et al. (10)). The order of magnitude of the conductance and the current histograms are very simi- lar. Equally similar traces were found for conduction bursts, flickering traces and other phenomena typical for protein con- ductance (10). Fig. 2 (bottom) a conduction burst in a synthetic membrane (top) is compared to a burst of the TRPM8 channel activity. The current histogram of the two bursts is very simi- lar, both in absolute currents and in the peak areas. It is again difficult to distinguish the data from the synthetic and cell mem- branes without independent experimental evidence. It is likely that this statement is generally true. 2.3 Pore geometries Lipid pores are transient in time and probably due to area fluc- tuations in the membrane. There exists no experimental evi- dence for a well-defined pore geometry and size from electron microscopy or other nanoscopic techniques. Glaser and col- laborators (12) proposed two kinds of pores: The hydrophobic pore shown in Fig. 3A (left) and the hydrophilic pore shown in Fig. 3A (right). The hydrophobic pore is a area density fluc- tuation without major rearrangement of its lipids. Water in the pore is in contact with hydrophobic hydrocarbon chains. The hydrophilic pore (Fig. 3A, right) is linked to a rearrangement of the lipids so that contact with water is avoided. Such pores are thought to be more stable and long-lived. Based on elastic Figure 1: Lipid ion channels in a synthetic membrane (DMPC:DLPC=10:1 in 150mM KCl, 30◦C). A. Charac- teristic current trace recorded at 50mV. It was stable for more than 30 minutes. B. The single-channel I-V profile is linear, resulting in a conductance of 330 pS. C. Life- time distribution of the channels. Adapted from Laub et al, 2012 (10). D. Ceramide channels in soybean lipids (asolectin) in 1M KCl. From Siskind et al., 2002 and 2003 (3, 11). fect on the physics of membranes, such drugs can 'block' lipid channels without binding to any particular receptor. Similarly, hydrostatic pressure shifts transitions upwards (4). It seems to be increasingly accepted that the composition of lipid membranes can influence the channel activity of pro- teins (7). Interestingly, critical phenomena such as the ones described above for synthetic membranes have also been found for protein channels embedded in synthetic lipid membranes. The characteristics of some of these channels are highly corre- lated with the chain melting transition in the surrounding mem- brane. For instance, the KcsA potassium channel when recon- stituted into a synthetic membrane displays a mean conduc- tance that exactly reflects the heat capacity profile of the mem- brane (8). Simultaneously, the open lifetime of the channel is maximum in the membrane transition. Thus, the properties of membranes containing this channel protein accurately reflect the physics of the fluctuations in the lipid membrane. Very similar phenomena were found for the sarcoplasmic reticulum calcium channel (9). 2 considerations, the pore diameter was estimated to be of or- der 1nm (12), suggesting a well-defined channel conductance. Both, hydrophilic and hydrophobic pores have been found in MD simulations (13) (Fig. 3B). Application of voltage across the membrane initially forms a hydrophobic pore which sub- sequently develops into a hydrophilic pore. Due to the lack of direct observation, these geometries must be regarded as ten- tative. The proposed structure of ceramide channels consists rather of a stable geometry with well-defined molecular order. Considering their longer lifetimes and larger conductance, it is not clear whether these events are comparable to the lipid ion channel events described in Fig. 1. Figure 2: Top: The comparison of channel events from synthetic lipid membranes (top) and from protein- containing cell membranes (center: TRPM2, bot- tom:TRPM8) demonstrate the phenomenological similar- ity of both stepwise conductance and probability distribu- tion of current events (10). Bottom: Conduction burst in a synthetic membrane (top) and in a HEK cell membrane containing the TRPM8 channel (bottom) (10). The cur- rent histograms are nearly identical. Figure 3: Hydrophobic and hydrophilic pores in lipid membranes. A. Schematic drawing with hydrophobic pore on the left and hydrophilic pore on the right. B. Molecular dynamics simulation of a membrane subject to a voltage difference of 2 V(13). Here, the the hy- drophobic pore (left) consists of a file of water molecules spanning through the membrane. Such pores are thought to be dynamic structures caused by fluctuations in the lipid bilayer. C. Hypothetical structure of a ceramide channel.(14). 3 Transitions in the membrane and per- meability maxima Lipid membranes display melting transitions. In these transi- tions the membranes remain intact, but enthalpy ∆H and en- tropy ∆S change at a defined temperature Tm = ∆H/∆S. This is schematically displayed in Fig. 4 (a, b). One can in- vestigate such transitions in calorimetry by recoding the heat capacity, cp = (dH/dT )p. The heat capacity profile shown in panel c is from synthetic vesicles of dipalmitoyl phosphatidyl- choline (the lipid in panel a) with a melting temperature of ∼41◦C. Panel d (shaded area) shows the cp-profile of native E. coli membranes with a lipid melting peak close to physi- ological temperatures (Tm ≈ 22◦C). This has also been re- ported for other bacterial membranes, for lung surfactant (4), and for nerve membranes from rat brains, where the heat ca- pacity maximum is always found in the range from 20-30◦C. It seems likely that chain melting close to physiological condi- tions is a generic property of cells. The integrated heat capacity yields both ∆H and the ∆S. Simultaneously, the volume of the membrane changes by about 4% and the area by about 24% (17). The fluctuation-dissipation theorem (FDT) states that heat capacity cp, volume and area compressibility κV T are given by T and κA cp = 3 (cid:10)H 2(cid:11) − (cid:104)H(cid:105)2 RT 2 ; Figure 4: Melting of lipid membranes and permeability changes. a, b. Schematic representation of the melting process in a lipid membrane (4). c. Calorimetric melting profile of a dipalmitoyl phosphatidylcholine (DPPC) membrane. d. Melting profile of native E. coli membranes (4). The peak shaded in red is the lipid melting peak, situated about 10-15 degrees below growth temperature (dashed line). e. Permeability of a synthetic lipid membrane for a fluorescence dye compared to its heat capacity (15). f. Conductance of a synthetic lipid membrane for ions compared to the heat capacity (16). (cid:10)V 2(cid:11) − (cid:104)V (cid:105)2 (cid:10)A2(cid:11) − (cid:104)A(cid:105)2 (cid:104)V (cid:105) RT (cid:104)A(cid:105) RT κV T = κA T = ; , (1) 3.1 Relaxation timescales and lipid channel life- times The lifetime of fluctuations is also described by the fluctuation- dissipation theorem (5). Generally, larger fluctuations are asso- ciated with larger timescales. In the case of single lipid mem- i.e., they are related to enthalpy, volume and area fluctuations. Empirically, it was found that T = γV · cp κV ; T = γA · cp , κA (2) where γV ≈ 7.8 · 10−10m2/N and γA ≈ 0.89 m/N are mate- rial constants (17 -- 19). As a consequence, membranes become very soft in transitions. The application of the FDT to lipid membranes is discussed in detail by Heimburg (2010) (1). Several previous studies reported that membranes become more permeable in the transition regime (1). In order to create a pore, work ∆W (a) has to be performed: ∆W (a) = 1 2κA T A0 a2 , (3) where a is the area of the pore and A0 is the area of the overall membrane. Here, the line tension of the pore circumference(12, 20) is not explicitly contained for reasons discussed in Blicher et al. T has a maximum in the transition, the likelihood of finding a pore is enhanced and the permeability is high. This is shown for fluorescence dyes and ions in Fig. 4 (e and f). (15). Since κA 4 Figure 5: Left: Relaxation time scale of DPPC MLV com- pared to the heat capacity (6). Right: Open lifetimes of channel events in a D15PC: DOPC = 95:5 mixture in the fluid phase and in the phase transition (16). branes close to transitions it has been shown that relaxation times are reasonably well described by an single-exponential process with a time constant of τ = T 2 L ∆cp , (4) where L is a phenomenological coefficient with L = 6.6 · 108 J K/mol s for multilamellar lipid vesicles (MLV) of synthetic lipids (6). In nonequilibrium thermodynamics it is assumed that the relaxation and the fluctuation timescales are identical. Therefore, the time scale τ is intimately related to the mean open time of lipid channels. Fig. 5 (left) shows that DPPC MLV display a ∆cp maximum of about 400 kJ/mol K, leading to a relaxation time τ of 45 seconds. The half width of this transition is of the order of 0.05-0.1 K. The half width of the heat capacity profile of lung surfactant is much broader (order 10 K). The heat capacity at maximum is about 300 times smaller than for DPPC, and one expects a max- imum relaxation time on the order of 100 ms. Similar numbers are expected for the transitions in E. coli and other bacterial membranes, and the membranes of nerves. Interestingly, 1 - 100 ms is the timescale of the open lifetime of protein channels -- as well as that of lipid channels. The right hand panel of Fig. 5 shows the distribution of lipid channel lifetimes within and above the melting range of a synthetic lipid mixture (16). It can be seen that the lifetime increases 5 to 10-fold in the tran- sition range, which is consistent with the above considerations. Below, we will show that the KcsA potassium channel displays similar dependence on the melting transition in the surrounding membrane. 3.2 Gating of lipid channels The likelihood of finding lipid pores depends on experimental conditions due to the dependence of the melting point on all intensive thermodynamic variables. The permeability must be considered as primarily due to pore formation. Therefore, the likelihood of finding channels is correlated with changes in the intensive variables. In analogy with the nomenclature of protein channels, we will call this ef- fect 'gating'. Gating implies that the open probability of lipid channels depends in a very general sense on the intensive ther- modynamic variables. It has been shown experimentally that lipid channels can be gated by (1) • temperature (temperature-sensing). • lateral pressure or tension (mechanosensitive-gating). • general anesthetics (gating by drugs). • calcium and pH , i.e., chemical potential differences of calcium and protons. • voltage (voltage-gating), discussed below. Note that these variables have also been reported to control protein channels, e.g., the temperature sensitive TRP channels (21), mechanosensitive channels (22), the effect of general anes- thetics on the nicotinic acetyl choline receptor (23), calcium channels (9), the pH-dependent (8) and voltage-gated (7) KcsA potassium channels. 3.3 The effect of voltage At suitable voltages, one can induce single channel events in the synthetic membrane (Fig. 6). In contrast to the conduc- tance of the overall membrane, the single channel conductance is constant and leads to a linear current-voltage relation. The open probability increases as a function of voltage. Figure 6: Voltage-gating in a DMPC:DLPC=10:1 mem- brane at 30◦C in 150mM KCl. Top: Current-traces at four voltages showing an increase of single-channel conductance with voltage and an increased likelihood of channel formation. Bottom, left: The corresponding lin- ear single-channel current-voltage relation indicating a single-channel conductance of γ=320 pS. Bottom, right: Open probability as a function of voltage. The phase diagram of a membrane as a function of volt- age was recently given by Heimburg (24). If one regards the membrane as a capacitor, one can calculate the force on the membrane due to electrostatic attraction. This force can reduce the thickness of the membrane. It can thus change the melting temperature and potentially create holes above a threshold volt- age (20, 25). The electrostatic force, F, exerted by voltage on a planar membrane is given by F = 1 2 CmV 2 m D (5) where Cm is the membrane capacitance, Vm is the transmem- brane voltage and D is the membrane thickness (24). This force reduces the thickness of the membrane (2). The electrical work performed on the membrane by a change in thickness from D1 to D2 is (cid:90) D2 ∆Wel = FdD ∝ V 2 m . (6) D1 5 It can thus be assumed that the free energy of pore forma- tion, ∆G, is related to the square of the voltage and to the elas- tic constants of the membrane (2) with ∆G = ∆G0 + αV 2 m , (7) where ∆G0 is the free energy difference between open and closed pores in the absence of voltage and α is a constant. ∆G0 reflects the elastic properties of the membrane that depend on composition, temperature and pressure. For asymmetric mem- branes one obtains ∆G = ∆G0 + α(Vm − V0)2 , (8) where the offset voltage V0 is due to membrane curvature or to a different lipid composition in the two membrane leaflets (27). The probability, Popen(Vm), of finding an open pore in the membrane at a fixed voltage is given by (cid:18) (cid:19) Popen(Vm) = K(Vm) 1 + K(Vm) ; K(Vm) = exp − ∆G kT , (9) where K(Vm) is the voltage-dependent equilibrium constant between open and closed states of a single pore. The current-voltage relation for the lipid membrane is pro- portional to the likelihood of finding an open channel for a given voltage: Im = γp · Popen · (Vm − E0) , (10) where γp is the conductance of a single pore and E0 = is the Nernst potential. While the voltage V0 reflects the asymmetry of the membrane, E0 reflects the asymmetry of the ion concen- trations of the buffer solution. If the aqueous buffer is the same on both sides of the membrane, the Nernst potential is zero. Fig. 7 (left) shows the I-V profile of a lipid membrane made of a mixture of DMPC and DLPC (10) and a fit given by the Figure 7: Current voltage-relations. Left: A synthetic membrane (DMPC:DLPC=10:1). The insert is the open probability of a membrane pore. The solid line repre- sents a fit to eq. 10 with E0 = 0 V. Right: I-V profiles of TRPM8 channels in HEK cells, adapted from Voets et al. 2004 (21) (top panel) and of TRPM5 in HEK cells, adapted from Talavera et al. 2005 (26). The solid lines are fits to eq. 10. 6 formalism given by eqs. 8 to 10 (V0 = −110mV, γm = 6.62 nS, ∆G0 = 5.2 kJ/mol, and a = −248 kJ/mol·V2). This fit reproduces the experimental profile. For comparison, the right hand panel of Fig. 7 shows the current-voltage relation- ships of two proteins from the TRP family. Members of this family of ion channels have been reported to respond to envi- ronmental stimuli such as temperature, membrane tension, pH and various drugs (28). No particular structure for these chan- nels is known, and no very well-defined selectivities for ions have been reported. Fig. 7 (right panels) show data for the current-voltage relationship of TRPM8 at two temperatures and TRPM5 adapted from publications of B. Nilius' group (21, 26). The I-V profiles look quite similar to that obtained from the synthetic membrane. The solid lines in Fig. 7 (right panels) are fits to the above formalism using parameters of similar order of magnitude as used for the synthetic membrane. The qual- ity of these fits indicates that the TRP channel conductance is well described as unspecific pore formation in an asymmetric membrane caused by the charging of the membrane capacitor. 4 The role of proteins Here, we consider several cases where proteins and lipid pores display similar dependences on intensive variables. 4.1 Temperature dependence of lipid pores, tem- perature sensing protein channels and van't Hoff law Sensitivity to temperature is one of the most prominent proper- ties of the TRP channels (26). Such channels display a temper- ature sensitivity over a temperature range of 10 K that is simi- lar to the width of melting transitions in many biomembranes. For instance, the TRPM8 channel is activated at temperatures about 10 degrees below body temperature, just where the max- imum of the melting profile of many cell membranes is found. While this may be coincidental, the assumption of temperature- sensing macromolecules is problematic as we discuss below. Assume a channel protein with open and closed states as shown in Fig. 8A. These states correspond to distinct pro- tein conformations. The equilibrium constant between the two states is K = exp (−∆G/RT ) with ∆G = ∆H − T ∆S. The likelihood for finding open and closed states is Popen = K 1 + K and Pclosed = 1 1 + K (11) respectively (van 't Hoff's law). On this basis, Talavera et al. reported activation enthalpies on the order of 200kJ/mol for TRPM4, TRPM5, TRPM8 and TRPV1 channels (26). Fig. 8B shows a fit of the conductance of a TRPM8 channel (21) to eq. 11. The corresponding transition enthalpy is ∆H =-250 kJ/mol, and the entropy ∆S =-856 J/mol·K. Fig. 8C displays the corresponding heat capacity of the transition in the protein with a maximum at 18.9 ◦C. It is given by the derivative of the fit in Fig. 8B multiplied with the van't Hoff enthalpy. These numbers are comparable to the total heat of protein unfolding. Typical values are: about 350kJ/mol for staphylococcal ribonu- clease, 200 kJ/mol for lysozyme, and 200-500 kJ/mol for met- myoglobin. However, such transition enthalpies seem highly Figure 8: A. Schematic drawing of a possible equilib- rium between an open and a closed state of a channel protein. The equilibrium is defined by Gibbs free energy, enthalpy and entropy differences (∆G, ∆H and ∆S, re- spectively). B. Van 't Hoff analysis of the temperature de- pendent conductance of a TRPM8 channel in HEK cells at -80mV (adapted from (21)). It leads to ∆H of -250 kJ/mol and ∆S=-859 J/mol· K for the two-state equilib- rium. C. From the analysis in panel B one obtains a heat capacity profile of the transition with a midpoint at 18.9 ◦C. D. For comparison, the lipid melting profile of native E. coli membranes is shown. It displays similar transition width and midpoint. unlikely for the small conformational change from an open to a closed state. The conformational change from closed to open state of a protein should have a much smaller enthalpy change and should thus have a much smaller temperature dependence. Similarly, the claimed difference in the entropy of closed and open states is too large. According to Boltzmann's equation, S = k ln Ω (with Ω being the degeneracy of states), an en- tropy difference of -856 J/mol K corresponds to a change of the number of states by a factor of Ω = 5 · 1044. A change of this magnitude is plausible for protein denaturation where there is no well-defined unfolded structure but one well defined native conformation. It is not reasonable for a transition between two states with well-defined function and geometry. Due to cooper- ative behavior, however, the melting of the lipid membrane can easily have enthalpies and entropies of the above order. Fig. 8D shows the experimental melting profile of E. coli membranes. The activation profile of TRPM8 channels is apparently quite consistent with the melting profile of a cell membrane. Figure 9: The dependence of KcsA channels on the phase transition of its host membrane. Top: The conductance of the channel in two lipid mixtures (symbols) compared to the respective heat capacity profile of the lipids (lines). Bottom: Open lifetimes of the KscA-channel (symbols) compared to the heat capacity profile (line). From Seeger et al. (2010) (8). 4.2 Channel proteins and phase transitions in the lipid membrane The KcsA potassium channels are pH- and voltage-gated. In or- der to measure their properties they are frequently reconstituted into synthetic membranes such as POPE:POPG=3:1 mol:mol (8). It is unclear why this particular lipid mixture is often cho- sen. Measurement of the heat capacity reveals that this lipid mixture has a melting profile with a maximum close to room temperature (blue line in Fig. 9, top). Interestingly, the mea- surement of the mean conductance of the KcsA channel recon- stituted into this membrane (blue symbols) exactly follows the heat capacity profile. This is not accidental since a change in the lipid composition to POPE:POPG=1:1 reveals that the con- ductance profile of the channel shifts in the same manner as the heat capacity curve (black symbols). A similar observation can be made for the KcsA channel open times. Typically, the open time distribution is fitted with a biexponential yielding two time constants. These two time constants are plotted in Fig. 9 (bot- tom) for the POPE:POPG=3:1 mixture as a function of temper- ature and compared with the respective heat capacity profiles. It was found that the lifetimes also follow the heat capacity pro- 7 Figure 10: The interaction of proteins with lipid membranes and hydrophobic matching. The top row refers to a protein that favors the solid state, while the bottom row refers to a protein favoring the liquid state. Left: Schematic drawing of the lipid arrangement around proteins. Center: The influence of proteins on the heat capacity profile of a synthetic membrane (Top: Band 3 protein of erythrocytes in DMPC shifts heat capacity profiles towards higher temperatures. Bottom: Cytochrome B5 in DPPC shifts cp profiles towards lower temperatures). Right: Monte Carlo simulations of the local fluctuations (yellow indicates larger fluctuations) of the lipid membrane close to the protein. Top: Fluctuations are enhanced at the protein interface above the melting temperature Tm and unaffected below. Bottom: Fluctuations are enhanced below Tm and unaffected above the transition. file. Very similar behavior was found for the sarcoplasmic retic- ulum calcium channel reconstituted into POPE:POPC mixtures at different ratios (9). Close to the largest heat capacity events of the lipid mixture, the channel displayed maximum activity (highest conductance) and the longest open times. The behavior described above is expected from the fluctu- ation dissipation theorem for the lipid transition itself. Both mean conductance and the lifetimes of conduction events ac- curately reflect the physics of the lipid membrane. Interest- ingly, in the case of the KcsA channel the conductance (and thus channel activity) is related to the total amount of protein in the membrane and is still inhibited by the potassium channel blocker tetraethyl ammonium (TEA). Channel conductance in these systems is apparently a property of the lipid-protein en- semble. In the following we suggest one possible description for this behavior. 4.3 A possible catalytic role of membrane pro- teins There is ample evidence in the literature that membrane pro- teins can influence the thermodynamic properties of lipid mem- branes. This is true for both integral and peripheral proteins. Fig. 10 (center, top) shows the calorimetric profiles of the band 3 protein of erythrocytes, and of cytochrome b5 (Fig. 10, center bottom) reconstituted into synthetic lipid membranes (29, 30). While band 3 protein increases the temperature regime of mem- brane melting in membranes, cytochrome bb lowers it. Similar observations have been made with other proteins. E. coli mem- branes display a lipid transition around 22 ◦C (Fig. 4 d) while 8 the extracted lipids (in the absence of proteins) display a tran- sition around 12 ◦C. Since proteins influence melting points, they must also affect lipid membrane fluctuations and the oc- currence of lipid channels. The influence of an integral protein on membrane melting is partially dictated by the so-called hydrophobic matching (31). If the the hydrophobic part of the protein is more extended than the hydrocarbon core of a fluid membrane, it will match bet- ter with ordered lipids. As a consequence, lipids tend to be more gel-like at the interface of the protein (32). If the proteins have short hydrophobic cores (e.g., gramicidin A), it will fa- vor fluid lipids in its vicinity (Fig. 10, left). The first class of proteins will shift melting events towards higher temperatures, while the second class will shift it towards lower temperatures. This is shown for the experimental examples Band 3 protein and cytochrome b5 (Fig. 10, center). Peripheral proteins can shift transitions, for instance by shielding electrostatic charges on the surface. If a protein that matches the gel state of the membrane is located in a fluid membrane, it tends to surround itself by a gel lipid layer. As a consequence, there is a regime of high fluctuations near the protein (32, 33) meaning that both heat capacity and compressibility can be altered close to a protein. For this reason, the presence of a protein has the potential to locally induce lipid pores in its proximity. Another way of stat- ing this is that proteins can catalyze lipid pores at their outer interface. Fig. 10 (right) show Monte Carlo simulations of such a simulation (32, 33). The simulations show a protein (black) in a membrane at temperatures below and above the melting regime. Dark red shades indicate small fluctuations, while bright yellow shades display large fluctuations. It can be seen that a protein that favors the fluid lipid state would tend to create regimes of large fluctuations in its environment at tem- peratures below the melting temperature (top panels). It thus seems likely that proteins can catalyze channel ac- tivity without being channels themselves. This is due to the effect of the proteins on the cooperative fluctuations in the lipid membrane. It should generally be possible to estimate this ef- fect from the influence of the protein on melting transitions. 5 Summary The aim of this review has been to characterize lipid ion chan- nels, and to illustrate the similarity of ion conduction events and protein channel activity. We have shown that the appear- ance of lipid channels is rooted in the fluctuation dissipation theorem. It is thus strictly coupled to the thermodynamics of the membrane and influenced by changes in the thermodynam- ics variables such as temperature, pressure, voltage, etc. Many (but not all) properties of protein channels are practi- cally indistinguishable from lipid channels. These include • Single channel conductances and lifetimes • The current-voltage relations of some proteins such as TRPM8 and the synthetic membrane • The activation of TRP channels by temperature • The conductance and the lifetimes of KcsA and calcium channels embedded in membranes with transitions. However, a few important properties are difficult to recon- cile with the pure lipid membrane: • The effect of mutations in the protein. • The action of strong poisons such as tetrodotoxin or tetra- ethyl ammonium. Tedrodotoxin in high concentrations (mM regime) only displays a very minor influence on the melting profiles of zwitterionic membranes (unpublished data from Master's thesis of S. B. Madsen, NBI 2012). • The selectivity of the ion conduction, for instance of the potassium channel (about 10000 times higher conduc- tance for potassium over sodium). Lipid channels seem to display a mild selectivity only, following the Hofmeis- ter sequence (34). Nevertheless, there exist a number of cases, where one can demonstrate clear correlations of protein behavior with the lipid membrane physics, in particular the KcsA channels and cal- cium channels. It seems likely that a view will eventually emerge in which the conductance of biomembrane is seen as a feature of a lipid- protein ensemble rather than as a feature of single proteins. This implies a strong coupling to the macroscopic thermody- namics of the biological membrane as a whole. Acknowledgement: Thanks to Andrew D. Jackson from the Niels Bohr International Academy for a critical reading of 9 the manuscript. This work was supported by the Villum Foun- dation (VKR022130). References 1. Heimburg, T. 2010. Lipid ion channels. Biophys. Chem. 150:2 -- 22. 2. Blicher, A., and T. Heimburg. 2013. Voltage-gated lipid ion chan- nels. arXiv 1209.3640 [physics.bio-ph]:1 -- 10. 3. Siskind, L. J., R. N. Kolesnick, and M. Colombini. 2002. Ce- ramide channels increase the permeability of the mitochondrial outer membrane to small proteins. J. Biol. Chem. 277:26796 -- 26803. 4. Heimburg, T., 2007. Thermal biophysics of membranes. Wiley VCH, Berlin, Germany. 5. Kubo, R. 1966. The fluctuation-dissipation theorem. Rep. Prog. Phys. 29:255 -- 284. 6. Seeger, H. M., M. L. Gudmundsson, and T. Heimburg. 2007. How anesthetics, neurotransmitters, and antibiotics influence the relax- ation processes in lipid membranes. J. Phys. Chem. B 111:13858 -- 13866. 7. Schmidt, D., Q.-X. Jiang, and R. MacKinnon. 2006. Phospho- lipids and the origin of cationic gating charges in voltage sensors. Nature 444:775 -- 779. 8. Seeger, H. M., A. Alessandrini, and P. Facci. 2010. KcsA redis- tribution upon lipid domain formation in supported lipid bilayers and its functional implications. Biophys. J. 98:371a. 9. Cannon, B., M. Hermansson, S. Gyorke, P. Somerharju, and J. A. Virtanen. 2003. Regulation of calcium channel activity by lipid domain formation in planar lipid bilayers. Biophys. J. 85:933 -- 942. 10. Laub, K. R., K. Witschas, A. Blicher, S. B. Madsen, A. Luckhoff, and T. Heimburg. 2012. Comparing ion conductance record- ings of synthetic lipid bilayers with cell membranes containing trp channels. Biochim. Biophys. Acta 1818:1 -- 12. 11. Siskind, L. J., A. Davoody, N. Lewin, S. Marshall, and M. Colom- bini. 2003. Enlargement and contracture of c2-ceramide channels. Biophys. J. 85:1560 -- 1575. 12. Glaser, R. W., S. L. Leikin, L. V. Chernomordik, V. F. Pas- tushenko, and A. I. Sokirko. 1988. Reversible breakdown of lipid bilayers: Formation and evolution of pores. Biochim. Biophys. Acta 940:275 -- 287. 13. Bockmann, R., R. de Groot, S. Kakorin, E. Neumann, and H. Grubmuller. 2008. Kinetics, statistics, and energetics of lipid membrane electroporation studied by molecular dynamics simu- lations. Biophys. J. 95:1837 -- 1850. 14. Samanta, S., J. Stiban, T. K. Maugel, and M. Colombini. 2011. Visualization of ceramide channels by transmission electron mi- croscopy. Biochim. Biophys. Acta 1808:1196 -- 1201. 15. Blicher, A., K. Wodzinska, M. Fidorra, M. Winterhalter, and T. Heimburg. 2009. The temperature dependence of lipid mem- brane permeability, its quantized nature, and the influence of anes- thetics. Biophys. J. 96:4581 -- 4591. 30. Freire, E., T. Markello, C. Rigell, and P. W. Holloway. 1983. Calorimetric and fluorescence characterization of interactions be- tween cytochrome b5 and phosphatidylcholine bilayers. Bio- chemistry 28:5634 -- 5643. 31. Mouritsen, O. G., and M. Bloom. 1984. Mattress model of lipid- protein interactions in membranes. Biophys. J. 46:141 -- 153. 32. Ivanova, V. P., I. M. Makarov, T. E. Schaffer, and T. Heim- burg. 2003. Analizing heat capacity profiles of peptide-containing membranes: Cluster formation of gramicidin A. Biophys. J. 84:2427 -- 2439. 33. Seeger, H., M. Fidorra, and T. Heimburg. 2005. Domain size and fluctuations at domain interfaces in lipid mixtures. Macromol. Symposia 219:85 -- 96. 34. Antonov, V. F., A. A. Anosov, V. P. Norik, and E. Y. Smirnova. 2005. Soft perforation of planar bilayer lipid membranes of dipalmitoylphosphatidylcholine at the temperature of the phase transition from the liquid crystalline to gel state. Eur. Biophys. J. 34:155 -- 162. 16. Wunderlich, B., C. Leirer, A. Idzko, U. F. Keyser, V. Myles, T. He- imburg, and M. Schneider. 2009. Phase state dependent current fluctuations in pure lipid membranes. Biophys. J. 96:4592 -- 4597. 17. Heimburg, T. 1998. Mechanical aspects of membrane thermody- namics. Estimation of the mechanical properties of lipid mem- branes close to the chain melting transition from calorimetry. Biochim. Biophys. Acta 1415:147 -- 162. 18. Ebel, H., P. Grabitz, and T. Heimburg. 2001. Enthalpy and vol- ume changes in lipid membranes. i. the proportionality of heat and volume changes in the lipid melting transition and its implication for the elastic constants. J. Phys. Chem. B 105:7353 -- 7360. 19. Pedersen, U. R., G. H. Peters, T. B. S. der, and J. C. Dyre. 2010. Correlated volume-energy fluctuations of phospholipid membranes: A simulation study. J. Phys. Chem. B 114:2124 -- 2130. 20. Winterhalter, M., and W. Helfrich. 1987. Effect of voltage on pores in membranes. Phys. Rev. A 36:5874 -- 5876. 21. Voets, T., G. Droogmans, U. Wissenbach, A. Janssens, V. Flock- erzi, and B. Nilius. 2004. The principle of temperature-dependent gating in cold- and heat-sensitive TRP channels. Nature 430:748 -- 754. 22. Suchyna, T. M., S. E. Tape, R. E. Koeppe II, O. S. Andersen, F. Sachs, and P. A. Gottlieb. 2004. Bilayer-dependent inhibition of mechanosensitive channels by neuroactive peptide enantiomers. Nature 430:235 -- 240. 23. Bradley, R. J., R. Sterz, and K. Peper. 1984. The effects of alco- hols and diols at the nicotinic acetylcholine receptor of the neuro- muscular junction. Brain Res. 295:101 -- 112. 24. Heimburg, T. 2012. The capacitance and electromechanical cou- pling of lipid membranes close to transitions. the effect of elec- trostriction. Biophys. J. 103:918 -- 929. 25. Crowley, J. M. 1973. Electrical breakdown of bimolecular lipid membranes as an electromechanical instability. Biophys. J. 13:711 -- 724. 26. Talavera, K., K. Yasumatsu, T. Voets, G. Droogmans, N. Shige- mura, Y. Ninomiya, R. F. Margolskee, and B. Nilius. 2005. Heat activation of TRPM5 underlies thermal sensitivity of sweet taste. Nature 438:1022 -- 1025. 27. Alvarez, O., and R. Latorre. 1978. Voltage-dependent capacitance in lipid bilayers made from monolayers. Biophys. J. 21:1 -- 17. 28. Wu, L.-J., T.-B. Sweet, and D. E. Clapham. 2010. Interna- tional union of basic and clinical pharmacology. LXXVI. Current progress in the mammalian TRP ion channel family. Pharmacol. Rev. 62:381 -- 404. 29. Morrow, M. R., J. H. Davis, F. J. Sharom, and M. P. Lamb. 1986. Studies of the interaction of human erythroyte band 3 with mem- brane lipids using deuterium nuclear magnetic resonance and dif- ferential scanning calorimetry. Biochim. Biophys. Acta 858:13 -- 20. 10
1605.01366
1
1605
2016-05-04T18:12:13
Nanoscopic Characterization of DNA within Hydrophobic Pores:Thermodynamics and Kinetics
[ "physics.bio-ph", "cond-mat.mes-hall", "physics.chem-ph" ]
The energetic and transport properties of a double-stranded DNA dodecamer encapsulated in hydrophobic carbon nanotubes are probed employing two limiting nanotube diameters, D=4 nm and D=3 nm, corresponding to (51,0) and (40,0) zig-zag topologies, respectively. It is observed that the thermodynamically spontaneous encapsulation in the 4 nm nanopore (40 kJ/mol) is annihilated when the solid diameter narrows down to 3 nm, and that the confined DNA termini directly contact the hydrophobic walls with no solvent slab in-between. During the initial moments after confinement (2-3 ns),the biomolecule translocates along the nano pore's inner volume according to Fick's law (t) with a self-diffusion coefficient D=1.713 x 10-9m2/s, after which molecular diffusion assumes a single-file type mechanism (t1/2). As expected, diffusion is anisotropic, with the pore main axis as the preferred direction, but an in-depth analysis shows that the instantaneous velocity probabilities are essentially identical along the x, y and z directions. The 3D velocity histogram shows a maximum probability located at v=30.8 m/s, twice the observed velocity for a single-stranded three nucleotide DNA encapsulated in comparable armchair geometries (v=16.7 m/s, D=1.36-1.89 nm). Because precise physiological conditions (310 K, [NaCl] = 134 mM) are employed throughout, the present study establishes a landmark for the development of next generation in vivo drug delivery technologies based on carbon nanotubes as encapsulation agents.
physics.bio-ph
physics
Nanoscopic Characterization of DNA within Hydrophobic Pores: Thermodynamics and Kinetics Fernando J.A.L. Cruz,1,2,* Juan J. de Pablo2,3 and José P.B. Mota1 1LAQV, REQUIMTE, Departamento de Química, Faculdade de Ciências e Tecnologia, Universidade Nova de Lisboa, Caparica, 2Department of Chemical and Biological Engineering, University of Wisconsin-Madison, Madison, Wisconsin, 53706, USA 3Institute of Molecular Engineering, University of Chicago, Chicago, Illinois, 60637, USA 2829-516, Portugal.E-mail: [email protected] It is that observed respectively. The energetic and transport properties of a double-stranded DNA dodecamer encapsulated in hydrophobic carbon nanotubes are probed employing two limiting nanotube diameters, D = 4 nm and D = 3 nm, corresponding to (51,0) and (40,0) zig-zag topologies, the thermodynamically spontaneous encapsulation in the 4 nm nanopore (ΔG ≈ –40 kJ/mol) is annihilated when the solid diameter narrows down to 3 nm, and that the confined DNA termini directly contact the hydrophobic walls with no solvent slab in-between. During the initial moments after confinement (t ≤ 2–3 ns), the biomolecule translocates along the nanopore's inner volume according to Fick's law (∼ t) with a self-diffusion coefficient D = 1.713⋅10–9 m2/s, after which molecular diffusion assumes a single-file type mechanism (∼ t1/2). As expected, diffusion is anisotropic, with the pore main axis as the preferred direction, but an in-depth analysis shows that the instantaneous velocity probabilities are essentially identical along the x, y and z directions. The 3D velocity histogram shows a maximum probability located at <v> = 30.8 m/s, twice the observed velocity for a single-stranded three nucleotide DNA encapsulated in comparable armchair geometries (<v> = 16.7 m/s, D = 1.36–1.89 nm). Because precise physiological conditions (310 K and [NaCl] = 134 mM) are employed throughout, the present study establishes a landmark for the development of next generation in vivo drug delivery technologies based on carbon nanotubes as encapsulation agents. 1. INTRODUCTION A plethora of applications currently envisage carbon nanotubes (CNTs) as next-generation encapsulation media for biological polymers, such as proteins and nucleic acids 1, 2. Owing to several appealing features, such as large surface areas, well-defined physico-chemical properties and the hydrophobic nature of their pristine structure, CNTs are considered ideal candidates to be used as nanopores for biomolecular confinement. Present day potential applications span different purposes and objectives, such as intracellular penetration via endocytosis and delivery of biological cargoes 3-5, ultrafast nucleotide sequencing 6-8 and gene and DNA delivery to cells 5, 9, 10. The remarkable experimental work by Geng et al. 5, 11 has shown that carbon nanotubes can spontaneously penetrate the lipid bilayer of a liposome, and the corresponding hybrid incorporated into live mammalian cells to act as a nanopore through which water, ions and DNA are delivered to the cellular interior. For an efficient and cost- effective industrial fabrication of SWCNT-based technology for DNA encapsulation/delivery, the interactions between the solid and the biomolecule need to be thoroughly understood in order to render the DNA/SWCNT device able to be used under physiological conditions, T = 310K and [NaCl] = 134 mM. Nonetheless, the energetics and dynamics of single- (ssDNA) and double-stranded DNA (dsDNA) encapsulation onto single-walled carbon nanotubes (SWCNTs) are virtually unexplored and the corresponding molecular level details remain rather obscure. Previous theoretical and experimental work with DNA and SWCNTs has fundamentally been focused on the solids' external volume, overlooking the possibility of molecular encapsulation 12-14; nevertheless, it is well known that the conformational properties of biopolymers under confinement are of crucial relevance in living organisms (e.g., DNA packaging in eukaryotic chromosomes, viral capsids). Unrealistic high temperature (400 K) studies revealed that depending on pore diameter, a small 8 nucleobase-long ssDNA strand can be spontaneously confined 15. However, there is a critical diameter of 1.08 nm 16, bellow which molecular confinement is inhibited by an energy barrier of ca. 130 kJ/mol, arising essentially from strong van der Waals repulsions 17. These findings were extended for intratubular confinement of a 2 nm long ssDNA onto a SWCNT mimic of the constriction region of an α-hemolysin channel 18. dsDNA confinement in carbon nanotubes remains utterly uncharted, for most of the earlier work has focused on temperatures remarkably distinct from the physiological value, thus preventing extrapolation of results to in vivo conditions. The pioneer work of Lau et al. 19 showed that a small dsDNA molecule (8 basepairs long), initially confined onto a D = 4 nm diameter nanotube, exhibits a root-mean squared displacement similar to the unconfined molecule, but that behaviour is drastically reduced as the nanotube narrows to D = 3 nm; Cruz et al. 20, 21 have already demonstrated that diffusion inside SWCNTs can exhibit deviations from the classical Fickian behaviour. The insertion of dsDNA onto multi-walled carbon nanotubes has been experimentally observed by STM/STS, TEM and Rahman techniques 22, 23. However, it seems to be a competing mechanism with the wrapping of the biomolecule around the nanotube external walls; Iijima's reported data failed to identify the relevant conditions upon which the confinement process is favoured, such as ionic strength of the media and temperature 22. In order to probe the thermodynamical spontaneity of encapsulation, we have adopted the well-known Dickerson dodecamer 24 as dsDNA model (Figure 1) and conducted a series of well- tempered metadynamics calculations 25 involving two distinct nanotubes, namely (51,0) and (40,0) with skeletal diameters of D = 4 nm and D = 3 nm, respectively; very importantly, the results were obtained under precise physiological conditions, and the media ionic strength maintained at 134 mM by employing a NaCl buffer. sequence 24 with dodecamer B-DNA Dickerson Figure 1 – Dickerson dsDNA dodecamer. Isovolumetric representation of the 5'- D(*CP*GP*CP*GP*AP*AP*TP*TP*CP*GP*CP*G)-3', length L = 3.8 nm and skeletal diameter D = 2 nm, highlighting the changing chemical nature along the double strand; nucleobase residues are coloured according to their chemical nature, namely blue (A), red (T), purple (G) and brown (C). Note that the whole DNA molecule is atomistically detailed in the calculations, and thus each individual atom has a corresponding partial electrostatic charge. Our work indeed shows that the dsDNA molecule, initially in a bulk solution, can become encapsulated onto a D = 4 nm SWCNT leading to a pronounced decrease of the whole system's Gibbs free-energy 26. The de Gennes blob theory for polymers has been extended by Jun et al. 27 to include the effect of cylindrical confinement upon the biopolymer free energy, and the latter decreases with an increase of nanotube diameter; because the blob description breaks down once the DNA persistence length (strong confinement), Dai et al. 28 recently extended Jun's formalism to dsDNA strongly confined in slit- pore geometries. In order these encapsulation the present work provides a full thermodynamical mapping of the associated Gibbs surface as well as a structural and kinetic analysis of dsDNA within cylindrical nanopores. The remainder of the manuscript is organized as follows: molecular models and methods are the diameter approaches to address some of issues, described in Section 2, followed by a discussion of the main results obtained (Section 3) and finally highlighting some conclusions and future lines of work (Section 4). 2. METHODOLOGY AND ALGORITHMS 2.1 Molecular Models Molecules are described using atomistically detailed force fields, including electrostatic charges in each atom. The dispersive interactions are calculated with the Lennard-Jones (12,6) potential, cross parameters between unlike particles determined by the classical Lorentz-Berthelot mixing rules, and electrostatic energies described by Coulomb's law. DNA is treated as a completely flexible entity within the framework of the AMBER99sb-ildn force-field 29, 30, the corresponding potential energies associated with bond stretching, U(r), and angle bending, U(θ), are calculated with harmonic potentials, whilst the dihedral energies are computed using Ryckaert- U ( ) ϕ = 5 ∑ ∑ nC [ ( cos ϕ − º180 ] ) = 0 dihedrals n the diameter of Bellemans functions, . To retain computational tractability, we have chosen the double- stranded B–DNA Dickerson dodecamer 24, exhibiting a pitch 31 of P ~ 3.4 nm obtained from an average of 10–10.5 base- pairs per turn over the entire helix 32, and with a double-strand end-to-end length of L ~ 3.8 nm measured between terminal (GC) base pairs (Fig.1); the A–DNA form has P ~ 2.6 nm corresponding to an average of 11 base-pairs per turn 32. Considering that the B–DNA backbone P atoms lie on a cylindrical surface, the double-strand corresponds to D ~ 2 nm 31. In spite of smaller in length than genomic DNA, the Dickerson dodecamer main structural features resemble those of genomic λ-bacteriophage DNA 33, namely in the radius of gyration and double-strand backbone diameter, Rg ≈ 0.7–1 nm and D ≈ 2 nm. The Na+ and Cl– ions are described with the parameterization of Aqvist and Dang 34 and the H2O molecules by the TIP3P force field of Jorgensen and co-workers 35. Large diameter (D ≈ 4 nm) SWCNTs have been recently prepared by Kobayashi et al. 36. In order to examine DNA confinement into such large, hollow nanostructures, two different diameter SWCNTs were adopted, both with zig-zag symmetry and length L = 8 nm; skeletal diameters, measured between carbon atoms on opposite sides of the wall, are D = 4 nm (51,0) and D = 3 nm (40,0). The solid walls are built up of hexagonally-arranged sp2 graphitic carbon atoms, with a C–C bond length 21, 37 of 1.42 Å, whose Lennard-Jones potential is given by Steele's parameterization (σ = 0.34 nm, ε = 28 K) 38. 2.2 Molecular Dynamics and Metadynamics Molecular dynamics (MD) simulations in the isothermal– isobaric ensemble (NpT) were performed using the Gromacs set of routines 39. Newton's equations of motion were integrated with a time step of 1 fs and using a Nosé-Hoover thermostat 40, 41 and a Parrinello-Rahman barostat 42 to maintain temperature and pressure at 310 K and 1 bar. A the interactions, and potential cut-off of 1.5 nm was employed for both the van der Waals and Coulombic long-range electrostatics were calculated with the particle-mesh Ewald method 43, 44 using cubic interpolation and a maximum Fourier grid spacing of 0.12 nm. Three-dimensional periodic boundary conditions were applied throughout the systems. scheme of The well-tempered metadynamics Barducci and Parrinello 25 was employed to obtain the free- energy landscape associated with the confinement mechanism. Briefly, the well-tempered algorithm biases Newton dynamics by adding a time-dependent Gaussian potential, V(ζ,t), to the total (unbiased) Hamiltonian, preventing the system from becoming permanently trapped in local energy minima and thus leading to a more efficient exploration of the phase space. The potential V(ζ,t) is a function of the so-called collective variables (or order parameters), ζ(q) = [ξ1(q), ξ2(q), …, ξn(q)], which in turn are related to the microscopic coordinates of the real system, q, according to V ( ζ ( ) Wtq , = ) ⎛ exp ⎜ ⎝ − t t t ' ≤ ∑ 0' = V ) ( ( ) ) ( t tq ', ' ζ T Δ ⎞ ⎟ ⎠ (1) ⎛ exp ⎜ ⎜ ⎝ ( ζ i ( ) q − n ∑ i 1 = 2 ( ) ) ) ( tq ' − ζ i 2 2 σ i ⎞ ⎟ ⎟ ⎠ in DNA z the of !" ξ2 = R free-energy two collective !" GC GC − R 12 1 where t is the simulation time, W=τGω is the height of a single Gaussian, τG is the time interval at which the contribution for the bias potential, V(ζ,t), is added, ω is the initial Gaussian height, ΔT is a parameter with dimensions of temperature, σi is the Gaussian width and n is the number of collective variables in the system; we have considered τG = 0.1 ps, ω = 0.1 kJ/mol, ΔT = 310 K and σ = 0.1 nm. The parameter ΔT determines the rate of decay for the height of the added Gaussian potentials and when ΔT → 0 the well-tempered scheme approaches an unbiased simulation. Because SWCNTs are primarily one- landscape was dimensional symmetric, constructed terms variables, !" − R !" ξ1 = R is the positional vector of the centre of mass of the biomolecule !" ( R ), projected along the z– axis, or of the terminal (GC) nucleobase pairs at the double- !" strand termini, ( R ). According to our definition of collective variables, ξ1 corresponds to the z-distance between the biomolecule and the nanopore centre and ξ2 can be length. The characteristic the nanotube and Dickerson dodecamer are, respectively, L = 8 nm and L = 3.8 nm, and nm therefore any value corresponds the biomolecule is completely encapsulated within the solid ) and of the nanotube ( R!" !" ) and ( R DNA L − to a DNA–SWCNT hybrid =Δ=ξ 1 1.22 in which the DNA end-to-end !" , where R interpreted as ( SWCNT L lengths of SWCNT z and SWCNT z DNA z GC 12 GC 1 L ) < boundaries; the threshold ξ1 > 5.9 nm obviously indicates the absence of confinement. At the end of the simulation, the three-dimensional free-energy surface is constructed by summing the accumulated time-dependent Gaussian potentials F ( , ζ t ) −= T T Δ+ T Δ V ( , ζ )t t ) 0 t →∂ according to . A discussion of the algorithm's convergence towards the correct free-energy profile is beyond the scope of this work and can be found elsewhere 25, 45; suffices to say that it in the long time limit ( V ζ , ∂ and therefore the well-tempered method leads to a converged free-energy surface. An alternative approach to obtain the time-independent free-energy surface ( F ,ζ at the final portion of the relies on integrating metadynamics run. The converged free-energy can thus be mathematically obtained from equation (2): )t F ( ) ζ −= 1 τ V t t tot ( V , ζ ∫ tot − τ t ) dt (2) where ttot is the total simulation time and τ is the time window over which averaging is performed. We have implemented a convergence analysis for each collective variable, ξ1 and ξ2, by splitting the last 40 ns of simulation time into τ = 4 ns windows, and confirming the convergence of the bias potential V(ζ, t). This test ensures that the Gibbs map of Figure 2 is a good estimator of the free-energy changes associated with molecular encapsulation. 2.3 Statistical Fittings of Velocity profiles Velocity data were histogram reweighted with a bin width of 0.002 nm/ps (Figure 6) and analysed in terms of the Gram- Charlier peak function (eqn. 3) or a sigmoidal logistic function type 1 (eqn. 4), where p(v) is the probability and v is the velocity measured along the corresponding direction. A 2 π 3 z = 2 2 z − e Hz ,3 − 1 4 ∑+ 3 = i a i i ! ( ) zH i ⎞ ⎟ ⎟ ⎠ 4 z 3 6 z − 3 + = ⎛ ⎜ ⎜ ⎝ 4 ( zyxv , , ) (3) a c ( vvk − ) ( ) ( )zvyvxv , ( ) , (4) ( ) vp = p 0 + w = v v − c w , H 3 z ( ) vp = 1 −+ e 3. RESULTS AND DISCUSSION We have adopted two macroscopic descriptors to construct the free-energy landscape, FE (ξ1) and FE (ξ2), relating the centres of mass of DNA and SWCNT projected along the nanopore main axis (ξ1) and the DNA end-to-end length (ξ2) (cf. Section 2.2). An inspection of the corresponding Gibbs map (Figure 2) clearly indicates the existence of five distinct located sequentially along free-energy minima, all the nanopore internal volume, ξ1 < 1.8 nm. Because they represent thermodynamically identical free-energy basins, DNA can translate freely within the nanopore, easily moving from one free-energy minimum to the next, while exploring a maximum probability configuration path between adjacent minima. However, in order to irreversibly escape from those deep free- energy valleys, FE(ξ1, ξ2) ~ –40 kJ/mol–1, DNA has to overcome large energetical barriers, rendering the exit process towards the bulk thermodynamically expensive. Furthermore, an inspection of Figure 2 reveals a position where the DNA double-strand located at ξ1 = 0.149 nm corresponds to the absolute free-energy minimum, highlighting the nanopore centre as the thermodynamically favoured region for the encapsulated molecule. The residual free-energy region located between – 0.1 <ξ1 (nm) < 0 is physically meaningless, for the system is always positioned in a phase space comprised max = 2.9 nm; the between ξ1 probability distributions associated with ξ1 and ξ2, obtained by histogram reweighting with a minute binwidth of 0.001 nm are recorded in Supplementary Information (Fig. SI1). A detailed convergence analysis of the well-tempered metadynamics technique is beyond the scope of the present work and has been discussed elsewhere 45. Suffices to say that in the present case the individual free-energy profiles recorded in Figure 2, obtained by 4 ns integration time windows of the free-energy with respect to ξ1, clearly indicate that the technique has converged to the correct Gibbs map; although not shown, a similar conclusion upholds regarding ξ2. min = 5.4 ⋅ 10–6 nm and ξ1 , and ξ2 is the , where R!" Figure 2. Top) Gibbs free-energy landscape of dsDNA@(51,0) SWCNT. The Gibbs surface is built in terms of two distinct macroscopic descriptors, ξ1 and ξ2; ξ1 is the distance between centres of mass of the dsDNA and nanotube projected along the nanopore's main axis, ξ1 = R!" ξ2 = R!" GC 1 − R!" GC 12 DNA z − R!" SWCNT z DNA z is the SWCNT z ) and of the absolute dsDNA end-to-end length, positional vector of the centre of mass of the biomolecule ( R→ nanotube ( R→ strand termini, ( R→ ). Note that the consecutive free-energy minima along ξ1 indicate that dsDNA is free to translate along the nanotube. The snapshots were taken at different time intervals corresponding to (ξ1, ξ2) nm: A) (0.149, 4.112), B) (0.621, 4.164), C) (1.306, 4.164) and D) (1.794, 4.115); H2O molecules and Na+ and Cl– ions have been omitted for clarity sake. Bottom) ξ1 convergence profiles. Each set of data was obtained by a 4 ξ1, ns ) or of the terminal (GC) nucleobase pairs at the double- GC 1 ) and ( R→ free-energy regarding time GC 12 of integration + 1 τ j 4 −= ) ∫ ns 4 V τ j FE ( ξ 1 ( ξ 1 , t ) dt , performed over the last 40 ns of a )1ξFE ( in the top figure is an accurate representation of simulation run of 70 ns length. Note that the overall profile depicted in black, obtained in ∆t = 40–70 ns, exactly overlaps with the final 4 ns time window (66–70 ns, dark grey). It is clear that the free-energy profiles converged to the corresponding minima after ca. 60 ns, with only minor contributions being added from this time onwards; the Gibbs landscape recorded the thermodynamical free-energy changes associated with the encapsulation process. The narrowness of the employed (40,0) topology, D = 3 nm, by contrast with the (51,0) SWCNT, seems to prevent confinement even over an observation time window as large as 0.1 µs. Instead of encapsulation, the DNA molecule's direct contact with the (40,0) nanopore leads to the occurrence of two intrinsically distinct situations: i) owing to the strong π–π stacking interactions of terminal (GC) pairs with the graphitic mesh, the molecule exoadsorbs and threads the external SWCNT surface according to a mechanism previously observed by Zhao and Johnson 12, or ii) the biomolecule becomes trapped at the nanopore entrance, and partial melting of the double-strand terminii occurs. Using a (20,20) SWCNT with D = 2.67 nm to probe the encapsulation of siRNA, Mogurampelly and Maiti 46 showed that, due to a large free- energy barrier located at the nanopore entrance, confinement was thermodynamically prohibited. In order to better understand the energetics of interaction between the encapsulated molecule and the solid walls, an analysis of the van der Waals interactions has been implemented and the corresponding findings graphically recorded in Figure 3. Because the DNA centre-of-mass corresponds to the central inner tract formed by four nucleobase pairs (AATT, Fig.1), it is clear that its' interaction with the graphitic walls is rather minimal compared to the energies obtained from tracts; regardless of the observation time window, the termini energies always correspond to more that 87% of the dispersive two outer (CGCG) the radial ( ∑∑ r δ − i j i distribution function, r ij ) calculating ( ) rg = ( NNVV r i the ) j ≠ , where V is the volume of the system, Vr is the volume of a spherical shell at distance r from each particle i, Ni is the number of particles i in the system, rij is the distance between particles i and j and the triangular brackets denote an ensemble average over the entire simulation time window 47, 48. This procedure was performed for several different rij, namely the dsDNA and the SWCNT centres of mass (c.o.m.), the DNA end-to-end length, and the distance between the terminal (GC) basepairs and the graphitic inner surface. The results recorded in Figure 4 indicate that, upon encapsulation, the DNA's most probable site for occupancy is the nanopore centre, as indicated by the sharp peak located at 0.38 nm in Figure 4A, in striking agreement with the free energy minimum of ξ1 = 0.149 nm (Figure 2). As previously mentioned, the biomolecule is relatively free to travel between adjacent free-energy minima along the ξ1 direction, being able to explore the complete nanopore inner volume as indicated by the 0.4 nm and 3.62 nm grey peaks of Figure 4B: bearing in mind that the (51,0) nanotube employed in the calculations has a length of L = 8 nm, the r = 0.4 nm signal is equivalent to the pore centre and the r = 3.62 nm peak corresponds to one of the two symmetrical pore termini, where L/2 = 4 nm. The DNA's end-to-end length radial distribution function, measured between opposite (GC) pairs in the double strand, and recorded as black bars in Figure 4B, is well described by a Gaussian statistics with an average peak maximum located at r = 3.92 nm highlighting the stability of DNA's B-form characteristic of the Dickerson dodecamer (Figure 1). Furthermore, owing to the strong π–π interactions between the bare (GC) termini and the sp2 graphitic surface 37, the former are in direct contact with the latter, corroborating the absence of any solvent (hydrophilic) slab between the graphitic walls and the encapsulated DNA, as indicated by the 0.35 nm signal in Figure 3C; it should be noted that the van der Waals diameter of a graphitic carbon atom corresponds to 0.34 nm. The (40,0) topology, with an effective diameter of Deff = (3 – 0.34) = 2.66 nm, becomes too narrow for the DNA molecule its hydrophobic moieties (nucleobases) to directly contact the walls, such as the case for the (51,0) solid (Deff = 3.66 nm). Therefore, confinement is annihilated in the (40,0) nanotube because it would lead to an hybrid whose hydrophilic skeleton (phosphates) would be in direct contact with the hydrophobic carbons on the walls. to coil and allow energy of interaction between the whole DNA molecule and the solid walls, ULJ, and, on average correspond to 97%. It is also very interesting to observe that the distance between centres of mass of both molecules is essentially due to translocation of the DNA molecule along the nanopore main axis, for the corresponding 3D distance is either symmetrical about 0 (pore centre) or coincident with its z component (Figure 3); this comes to show that, in spite of the radial symmetry characteristic of nanotubes, the observed molecular mobility within the (51,0) encapsulating volume arises essentially from translation along the nanopore main axis and is severely restricted along the radial direction. Also, and apart from an initial time window of 20 – 22 ns, when the DNA molecule is accommodating itself to the confining volume, and when both energy and distance are considered at the same time, the constancy of the former with the latter demonstrates that the position of the DNA c.o.m. along the main axis does not determines the energy of interaction between biomolecule and solid, e.g., either located at the nanopore centre (d = 0 nm) or close to the termini ( d = 2 – 2.5 nm) the van der Waals energy is approximately constant. Figure 3. Dispersive energies and distances between dsDNA and (51,0) SWCNT. Energies are plotted their absolute values, U LJ ( ) ( )ϕθ U the biomolecule and the solid (black) and the energy term corresponding to the two (CGCG) tracts located at the double strand termini (blue). Distances interaction between UrU terms of ( ) total the for in + + = ( d x, y, z ) = d x( )2 + d y( )2 + d z( )2 a of has sequence correspond to the 3D distance, , between c.o.m. (black) and its z-component, d(z) (blue). The DNA double- strand 5'- D(*CP*GP*CP*GP*AP*AP*TP*TP*CP*GP*CP*G)-3', and thus the total van der Waals energy plotted in black corresponds to the sum of two terms: obtained from the two (CGCG) tracts, plotted in blue, and from the central (AATT) tract, not shown. To accurately characterize the structure and dynamics of confined dsDNA within the hydrophobic (51,0) nanopore, independent calculations have been conducted with the metadynamics biasing potential turned off, rendering the simulations exactly equivalent to Newton's law. The local structure associated with the biomolecule was retrieved by = − [ ( ) tr MSD D = lim t→∞ 1 6t MSD functions (Green-Kubo equation), and concluded that both approaches led to self-consistent results within the calculation error (0.5–1.0%). A similar conclusion was obtained by 53 for Takeuchi and Okazaki the self-diffusivities of polymethylene with O2 as a solute. Transport data were analysed herein by calculating ]20r ( ) the fluid mean-squared displacement, , and relating it with the self-diffusion coefficient using the , where r(t) is the Stokes-Einstein equation, positional vector of a unique molecule at time t, and the triangular brackets denote an ensemble average. Special care was taken to accurately sample the MD data; in order to achieve statistically significant results, we used small time delays between origins separated by 1 ps. A further analysis was conducted by decomposing the MSD obtained for the DNA center of mass into its x, y and z components. The results recorded in Figure 5A indicate that the biomolecule maintains its translational mobility whilst confined within the SWCNT, visiting a region of space whose boundaries are delimited by the nanopore termini, and associated with the adjacent free- energy minima identified in Figure 2 (ξ1 < 1.8 nm). It is very interesting to observe that during an initial time-window of ca. 2–3 ns the molecule diffuses according to Fick's law, e.g., the c.o.m. mean-squared displacement is linearly proportional to time, which corresponds to an effective self-diffusion coefficient of D = 1.789 ⋅ 10–9 m2/s. After that initial time- window, DNA approaches a single-file type diffusion 20, when the MSD becomes proportional to t1/2 ; translation within the nanotube is anisotropic. Indeed, the relative occurrences plotted that molecular displacement along the z direction (nanotube main axis) is favoured over the x and y analogues, corresponding to more than 33 % of the overall (3D) mean-squared displacement. This is clearly due to entropic reasons, for we have shown that the DNA molecule maintains direct contact with the solid walls (Fig. 4C) rendering the z direction as the preferred degree of freedom. in Figure 5B strongly indicate ( r 2 ⎤ σ = − ⎡ ⎢⎣ 0 r − A 1 A 2 ) 2 + ) + = φ 21 rr exp rr 1 ( exp − 0 ( )rg ( exp − ) ⎥⎦ Figure 4. Radial distribution functions at the (51,0) SWCNT. A) between DNA and SWCNT c.o.m., B) between (r1r24) and SWCNT c.o.m. (light grey) and (r12r13) and SWCNT c.o.m. (dark grey) and between (r1r24) and (r12r13) c.o.m. (black), equivalent to the DNA end-to-end length, C) between (r1r24) and SWCNT surface (light grey) and (r12r13) and SWCNT surface (dark grey). The blue lines correspond to the best statistical fittings of data with ( ) rg (A1 = 479.44, r1 = 0.53 nm, A2 = 68589.62, r2 = 0.10 nm, g0(r)= –1.80) (Figure A) and ( ) rg to eject ssDNA (ϕ = 24.40, r0 = 3.92 nm, σ = 0.15nm) (Figure B). Snapshots were taken to illustrate the local physical environment: ochre) DNA individual strands, grey) sp2 graphitic mesh and red) H2O oxygen atoms. For any nanotube-based drug delivery technology to find its way into the industrial production line, not only does the encapsulation mechanism needs to be thermodynamically favourable, therefore reducing energy costs, but it also needs to be reversible once the confined genetic material is ready to be delivered to the host cell. The recent theoretical work by Xue et al. 49 has demonstrated the feasibility of such ejection process, using filler agents (C60) and mechanical actuators (Ag nanowires) from within purely hydrophobic SWCNTs. Between encapsulation and ejection, the nucleic acid needs to travel from entrance to exit, corresponding to opposite nanopore entrances, in order to become available for cellular delivery: what happens in between? That is, how do we characterize the DNA kinetics once it is confined? One of the most relevant transport properties for industrial applications is the self-diffusion coefficient, D, which provides a measure of how mobile a fluid can be once it becomes encapsulated. This property can be determined from a time dependent analysis of molecular trajectories, using either the Stokes-Einstein or the Green-Kubo equations, which have become the de facto method 48, 50; both formalisms are mathematically equivalent 51. Recently, Baidakov et al. 52 determined the self-diffusion coefficient of stable and meta- stable Lennard-Jones fluids via the particles mean-squared displacement (Einstein equation) and velocity auto-correlation t to t MSD ∝ (Fick) and Figure 5. Kinetics of DNA confined at the (51,0) SWCNT. A) Mean- squared displacement profiles, MSD, for the molecular center of mass, and B) relative MSD occurrence regarding dimension i, MSDi MSD3D the classical diffusion mechanisms, lines correspond . The grey 2t MSD ∝ MSD ∝ (single-file) and the dashed (ballistic), black lines are the boundaries of the 3 – 4 ns time window after which molecular diffusion changes from Fickian to single-file type. black) Three- dimensional MSD3D, blue) MSD z-component, green) MSD x-component and purple) MSD y-component. Notice the anisotropic diffusion of DNA along the nanotube main axis, z, responsible for more than 33% of the effective 3D diffusion coefficient. In order to produce a nanoscopic picture of the DNA's mobility mechanism, the 3D instant c.o.m. velocity was calculated and the corresponding components recorded in Figure 6 after histogram reweighting with a bin width of 0.002 nm/ps. It now becomes clear that, although molecular mobility is anisotropic, favouring DNA the nanopore main axis, it is mostly due to entropic reasons, because identical independently of the particular direction in space. The in vitro experiments of Geng et al. 5 showed that porin-embedded nanotubes are able to transport ssDNA at an average speed of 1.5 nucleotides per millisecond towards the cellular internal volume; on the other hand, Pei and Gao 16 studied the translocation of a small three oligonucleotide ssDNA through armchair geometries, D = 1.36 nm and D = 1.89 nm, and observed an average translocation velocity of <v> = 16.66 m/s. Considering that the overall 3D velocity observed for the dsDNA c.o.m. is reasonably described by Gaussian statistics (Figure 6), a distribution maximum is observed at <v>max = 30.8 m/s. the kinetic energies translocation along (velocities) are Figure 6. Velocity probability profiles of confined dsDNA. Instant velocities were calculated for the entire observation time window (80 ns) and histogram reweighted with a bin width of 0.002 nm/ps: black) molecular center of mass (3D), blue) z-component, green) x-component and purple) y- component. Symbols correspond to histogram data and lines are the best statistical fittings (cf. Section 2.3); analysis along any individual direction (x, y, z) includes only the decay region. (Fig.2), and results resulting essentially favourable The encapsulation of an atomically detailed DNA dodecamer onto pristine (51,0) carbon nanotubes, D = 4 nm, is thermodynamically in DNA@SWCNT hybrids with lower Gibbs free-energy than the unconfined molecule (∆Gibbs ~ – 40 kJ/mol). Nonetheless, when the nanotube diameter narrows down to a (40,0) topology, D = 3 nm, encapsulation is completely inhibited. In this case, the biomolecule either gets stuck at the nanopore entrance or threads the solid outer walls in contact with the solid, as a direct consequence of strong π – π stacking between a terminal (GC) nucleobase pair and the graphitic mesh. Very interestingly, the dispersive energies between DNA and the SWCNT are rather independent of the former's position within the confining volume (Fig.3), and essentially result from the (CGCG) tracts located at the termini, e.g., more than 87%. The confined DNA maintains its translational mobility within the pore, <v>max = 30.8 m/s, however, due to entropic reasons, translocation is highly anisotropic from molecular displacement along the nanotube main axis, as verified by the decomposition between centres of mass distance (Fig.3) and mean-squared displacement (Fig.5) into their corresponding z components. This molecular translation occurs between adjacent free-energy minima, all located within the pore inner volume, and corresponding to a DNA conformation similar to the canonical B form. An inspection of the Gibbs free-energy map obtained for indicates a distance between consecutive free-energy minima of 0.47–0.69 nm. This has been overlooked before and seems to suggest the existence of a characteristical is unrelated to the graphitic mesh geometry; keeping in the mind the sp2 bond between Carbon atoms (0.142 nm), the hexagons composing the solid lattice are roughly 0.38 nm wide. To address this issue, further calculations would be compelling, employing much longer nanotubes and DNA strands. In order to keep tractable and atomistically detailed, a smaller system than the one being reported now could be employed, either by reducing the simulation box size or decreasing the outer walls solvation slab to focus on the endohedral volume. the calculations computationally length, which, however, we the (51,0) topology ACKNOWLEDGEMENTS This work makes use of results produced with the support of the Portuguese National Grid Initiative (https://wiki.ncg.ingrid.pt). F.J.A.L. Cruz gratefully acknowledges financial support from FCT/MCTES (Portugal) through grant EXCL/QEQ-PRS/0308/2012. think 4. CONCLUSIONS AND OUTLOOK REFERENCES 1. Kumar, H.; Lansac, Y.; Glaser, M. A.; Maiti, P. K., Biopolymers in Nanopores: Challenges and Opportunities. Soft Matter 2011, 7, 5898–5907. 2. Vashist, S. K.; Zheng, D.; Pastorin, G.; Al-Rubeaan, K.; Luong, J. H. T.; Sheu, F.-S., Delivery of Drugs and Biomolecules using Carbon Nanotubes. Carbon 2011, 49, 4077–4097. 3. Canton, I.; Battaglia, G., Endocytosis at the Nanoscale. Chem. Soc. Rev. 2012, 41, 2718–2739. 4. Kostarelos, K.; Bianco, A.; Prato, M., Promises, Facts and Challenges for Carbon Nanotubes in Imaging and Therapeutics. Nature Nano. 2009, 4, 627- 633. 5. Geng, J.; Kim, K.; Zhang, J.; Escalada, A.; Tunuguntla, R.; Comolli, L. R.; Allen, F. I.; Shnyrova, A. V.; Cho, K. R.; Munoz, D.; Wang, Y. M.; Grigoropoulos, C. P.; Ajo-Franklin, C. M.; Frolov, V. A.; Noy, A., Stochastic transport through carbon nanotubes in lipid bilayers and live cell membranes. Nature 2014, 514, 612–615. 6. Venkatesan, B. M.; Bashir, R., Nanopore Sensors for Nucleic Acid Analysis. Nature Nano. 2011, 6, 615-624. 7. Liu, H.; He, J.; Tang, J.; Liu, H.; Pang, P.; Cao, D.; Krstic, P.; Joseph, S.; Lindsay, S.; Nuckolls, C., Translocation of Single-Stranded DNA Through Single-Walled Carbon Nanotubes Science 2010, 327, 64-67. 8. Meng, S.; Maragakis, P.; Papaloukas, C.; Kaxiras, E., DNA Nucleoside Interaction and Identification with Carbon Nanotubes. Nano Letters 2007, 7, 45-50. 9. Wu, Y.; Phillips, J. A.; Liu, H.; Yang, R.; Tan, W., Carbon Nanotubes Protect DNA Strands During Cellular Delivery. ACS Nano 2008, 2, 2023– 2028. 10. Yum, K.; Wang, N.; Yu, M.-F., Nanoneedle: A multifunctional tool for biological studies in living cells. Nanoscale 2010, 2, 363–372. 11. Kim, K.; Geng, J.; Tunuguntla, R.; Comolli, L. R.; Grigoropoulos, C. P.; Ajo-Franklin, C. M.; Noy, A., Osmotically-Driven Transport in Carbon Nanotube Porins. Nano Lett. 2014, 14, 7051-7056. 12. Zhao, X.; Johnson, J. K., Simulation of Adsorption of DNA on Carbon Nanotubes. J. Am. Chem. Soc. 2007, 129, 10438-10445. 13. Alegret, N.; Santos, E.; Rodriguez-Fortea, A.; Rius, F. X.; Poblet, J. M., Disruption of small double stranded DNA molecules on carbon nanotubes: A molecular dynamics study. Chem. Phys. Lett. 2012, 525-26, 120-124. 14. Santosh, M.; Panigrahi, S.; Bhattacharyya, D.; Sood, A. K., Unzipping and Binding of Small Interfering RNA with Single Walled Carbon Nanotube: A platform for Small Interfering RNA Delivery. J. Chem. Phys. 2012, 136, 65106. 15. Gao, H.; Kong, Y.; Cui, D., Spontaneous Insertion of DNA Oligonucleotides into Carbon Nanotubes. Nano Letters 2003, 3, 471-473. 16. Pei, Q. X.; Lim, C. G.; Cheng, Y.; Gao, H., Molecular Dynamics Study on DNA Oligonucleotide Translocation Through Carbon Nanotubes. J. Chem. Phys. 2008, 129, 125101. 17. Lim, M. C. G.; Zhong, Z. W., Effects of Fluid Flow on the Oligonucleotide Folding in Single-walled Carbon Nanotubes. Phys. Rev. E 2009, 80, 041915. 18. Zimmerli, U.; Koumoutsakos, P., Simulations of Electrophoretic RNA Transport Through Transmembrane Carbon Nanotubes. Biophys. J. 2008, 94, 2546–2557. 19. Lau, E. Y.; Lightstone, F. C.; Colvin, M. E., Dynamics of DNA Encapsulated in a Hydrophobic Nanotube. Chem. Phys. Letters 2005, 412, 82–87. 20. Cruz, F. J. A. L.; Müller, E. A., Behavior of Ethylene/Ethane Binary Mixtures within Single-Walled Carbon Nanotubes. 2- Dynamical Properties. Adsorption 2009, 15, 13-22. 21. Cruz, F. J. A. L.; Müller, E. A.; Mota, J. P. B., The Role of the Intermolecular Potential on the Dynamics of Ethylene Confined in Cylindrical Nanopores. RSC Advances 2011, 1, 270-281. 22. Iijima, M.; Watabe, T.; Ishii, S.; Koshio, A.; Yamaguchi, T.; Bandow, S.; Iijima, S.; Suzuki, K.; Maruyama, Y., Fabrication and STM-characterization of Novel Hybrid Materials of DNA/carbon nanotube. Chem. Phys. Lett. 2005, 414, 520. 23. Ghosh, S.; Dutta, S.; Gomes, E.; Carroll, D.; Ralph D'Agostino, J.; John Olson; Guthold, M.; Gmeiner, W. H., Increased Heating Efficiency and Selective Thermal Ablation of Malignant Tissue with DNA-Encased Multiwalled Carbon Nanotubes. ACS Nano 2009, 3, 2667–2673. 24. Drew, H. R.; Wing, R. M.; Takano, T.; Broka, C.; Tanaka, S.; Itakura, K.; Dickerson, R. E., Structure of a B-DNA Dodecamer: Conformation and Dynamics. Proc. Nat. Acad. Sci. 1981, 78, 2179-2183. 25. Barducci, A.; Bussi, G.; Parrinello, M., Well-Tempered Metadynamics: A Smoothly Converging and Tunable Free-Energy Method. Phys. Rev. Lett. 2008, 100, 020603. 26. Cruz, F. J. A. L.; de Pablo, J. J.; Mota, J. P. B., Endohedral Confinement of a DNA Dodecamer onto Pristine Carbon Nanotubes and the Stability of the Canonical B Form. J. Chem. Phys. 2014, 140, 225103. 27.Jun, S.; Thirumalai, D.; Ha, B.-Y., Compression and Stretching of a Self- Avoiding Chain in Cylindrical Nanopores. Phys. Rev. Let. 2008, 101, 138101. 28. Dai, L.; Jones, J. J.; Maarel, J. R. C. v. d.; Doyle, P. S., A systematic study of DNA conformation in slitlike confinement. Soft Matter 2012, 8, 2972-2982. 29. Wang, J.; Cieplak, P.; Kollman, P. A., How Well Does a Restrained Electrostatic Potential (RESP) Model Perform in Calculating Conformational Energies of Organic and Biological Molecules? J. Comput. Chem. 2000, 21, 1049–1074. 30. Lindorff-Larsen, K.; Piana, S.; Palmo, K.; Maragakis, P.; Klepeis, J. L.; Dror, R. O.; Shaw, D. E., Improved side-chain torsion potentials for the Amber ff99SB protein force field. Proteins 2010, 78, 1950-1958. 31. Franklin, R. E.; Gosling, R. G., Molecular Configuration in Sodium Thymonucleate. Nature (London) 1953, 171, 740-741. 32. Vargason, J. M.; Henderson, K.; Ho, P. S., A crystallographic map of the transition from B-DNA to A-DNA. Proc. Nat. Acad. Sci. 2001, 98, 7265– 7270. 33. Wang, Y.; Tree, D. R.; Dorfman, K. D., Simulation of DNA Extension in Nanochannels. Macromolecules 2011, 44, 6594–6604. 34. Noy, A.; Soteras, I.; Luque, F. J.; Orozco, M., The Impact of Monovalent Ion Force Field Model in Nucleic Acids Simulations. Phys. Chem. Chem. Phys. 2009, 11, 10596–10607. 35. Jorgensen, W. L.; Chandrasekhar, J.; Madura, J. D.; Impey, R. W.; Klein, M. L., Comparison of Simple Potential Functions for Simulating Liquid Water. J. Chem. Phys. 1983, 79, 926-935. 36. Kobayashi, K.; Kitaura, R.; Nishimura, F.; Yoshikawa, H.; Awaga, K.; Shinohara, H., Growth of Large-diameter ((cid:0)4 nm) Single-wall Carbon Nanotubes in the Nanospace of Mesoporous Material SBA-15. Carbon 2011, 49, 5173–5179. 37. Cruz, F. J. A. L.; de Pablo, J. J.; Mota, J. P. B., Free Energy Landscapes of the Encapsulation Mechanism of DNA Nucleobases onto Carbon Nanotubes. RSC Advances 2014, 4, 1310-1321. 38. Steele, W. A., Molecular Interactions for Physical Adsorption. Chem. Rev. 1993, 93, 2355-2378. 39. Hess, B.; Kutzner, C.; Spoel, D. v. d.; Lindahl, E., GROMACS 4: Algorithms for Highly Efficient, Load-Balanced, and Scalable Molecular Simulation. J. Chem. Theory Comp. 2008, 4, 435–447. 40. Nosé, S., A unified formulation of the constant temperature molecular dynamics methods. J. Chem. Phys. 1984, 81, 511-519. 41. Hoover, W. G., Canonical Dynamics: Equilibrium Phase-space Distributions. Phys. Rev. A 1985, 31, 1695-1697. 42. Parrinello, M.; Rahman, A., Polymorphic Transitions in Single Crystals: A New Molecular Dynamics Method. J. Appl. Phys. 1981, 52, 7182-7190. 43. Darden, T.; York, D.; Pedersen, L., Particle Mesh Ewald: An N(cid:0)log(N) Method for Ewald Sums in Large Systems. J. Chem. Phys. 1993, 98, 10089- 10092. 44. Essmann, U.; Perera, L.; Berkowitz, M. L.; Darden, T.; Lee, H.; Pedersen, L. G., A Smooth Particle Mesh Ewald Potential. J. Chem. Phys. 1995, 103, 8577-8592. 45. Laio, A.; Gervasio, F. L., Metadynamics: a Method to Simulate Rare Events and Reconstruct the Free Energy in Biophysics, Chemistry and Material Science. Rep. Prog. Phys. 2008, 71, 126601. 46. Mogurampelly, S.; Maiti, P. K., Translocation and Encapsulation of siRNA Inside Carbon Nanotubes. J. Chem. Phys. 2013, 138, 034901. 47. Rowlinson, J. S.; Swinton, F. L., Liquids and Liquid Mixtures. Butterworths: London, 1982. 48. Haile, J. M., Molecular Dynamics Simulation: Elementary Methods. Wiley: New York, 1992. 49. Xue, Q.; Jing, N.; Chu, L.; Ling, C.; Zhang, H., Release of encapsulated molecules from carbon nanotubes using a displacing method: a MD simulation study. RSC Adv. 2012, 2, 6913–6920. 50. Allen, M. P.; Tildesley, D. J., Computer Simulation of Liquids. Clarendon Press: Oxford, 1990. 51. Zwanzig, R., Time-correlation functions and Transport Coefficients in Statistical Mechanics. Annu. Rev. Phys. Chem. 1965, 16, 67-102. 52. Baidakov, V. G.; Kozlova, Z. R., The self-diffusion Coefficient in Metastable States of a Lennard–Jones Fluid. Chem. Phys. Letters 2010, 500, 23–27. 53. Takeuchi, H.; Okazaki, K., Molecular Dynamics Simulation of Diffusion of Simple Gas Molecules in a Short Chain Polymer. J. Chem.Phys. 1990, 92, 5643-5652.
1704.08320
1
1704
2017-04-26T19:38:28
AFM-Assisted Fabrication of Thiol SAM Pattern with Alternating Quantified Surface Potential
[ "physics.bio-ph" ]
Thiol self assembled monolayers (SAMs) are widely used in many nano- and bio-technology applications. We report a new approach to create and characterize a thiol SAMs micropattern with alternating charges on a flat goldcoated substrate using atomic force microscopy (AFM) and Kelvin probe force microscopy (KPFM). We produced SAMs patterns made of alternating positively charged, negatively charged, and hydrophobic terminated thiols by an automated AFM assisted manipulation, or nanografting. We show that these thiol patterns possess only small topographical differences as revealed by AFM, and distinguished differences in surface potential (20 to 50 mV), revealed by KPFM. The pattern can be helpful in the development of biosensor technologies, specifically for selective binding of biomolecules based on charge and hydrophobicity, and serve as a model for creating surfaces with quantified alternating surface potential distribution.
physics.bio-ph
physics
Moores et al. Nanoscale Research Letters 2011, 6:185 http://www.nanoscalereslett.com/content/6/1/185 NANO EXPRESS AFM-assisted fabrication of thiol SAM pattern with alternating quantified surface potential Bradley Moores1, Janet Simons2, Song Xu3, Zoya Leonenko1,2* Open Access Abstract Thiol self-assembled monolayers (SAMs) are widely used in many nano- and bio-technology applications. We report a new approach to create and characterize a thiol SAMs micropattern with alternating charges on a flat gold- coated substrate using atomic force microscopy (AFM) and Kelvin probe force microscopy (KPFM). We produced SAMs-patterns made of alternating positively charged, negatively charged, and hydrophobic-terminated thiols by an automated AFM-assisted manipulation, or nanografting. We show that these thiol patterns possess only small topographical differences as revealed by AFM, and distinguished differences in surface potential (20-50 mV), revealed by KPFM. The pattern can be helpful in the development of biosensor technologies, specifically for selective binding of biomolecules based on charge and hydrophobicity, and serve as a model for creating surfaces with quantified alternating surface potential distribution. Background Thiol self-assembled monolayers (SAMs) are promising for many nano- and bio-technology applications as they offer a reliable method to produce surfaces with desir- able properties. These properties can be used for specific and non-specific binding of biomolecules and nanoparti- cles and, therefore, can serve as useful templates for nano- and micro-fabrication. The first systematic study of thiol chemicals was reported by Zisman and co- authors [1], and has since been investigated by many researchers, including a detailed review by Chechik et al [2]. SAMs can be defined as "molecular assemblies that are formed spontaneously be the immersion of an appropriate substrate into a solution of an active surfac- tant in an organic solvent" [3]. Thiols are a perfect type of such surfactant as they consist of a surface-active sul- fur group that binds to the metal surface, a hydrocarbon chain of various lengths that defines the packing of the monolayer, and a functional group at the end that deter- mines the functional properties of the formed SAM film. When metallic surfaces such as gold, platinum, or silver are exposed to thiols dissolved in organic solvent, a bond is formed between the thiol's active sulfur group and metal atoms of the surface, which is characterized * Correspondence: [email protected] 1Department of Physics and Astronomy, University of Waterloo, 200 University Avenue West, Waterloo, ON N2L 3G1, Canada. Full list of author information is available at the end of the article by a shared pair of electrons. Uniform monolayer cover- age can be created on flat metallic surfaces using proce- dures empirically determined for each thiol type, involving factors such as incubation time, solvent, and concentration [4,5]. With the invention of scanning probe microscopy and other nanoscale characterization techniques, much interest has been created in nanoscale fabrication for nanoelectro- nics and biosensing. The advantage of sensing on the nanoscale using miniaturized devices created a demand in producing thiol SAMs of a repeated pattern, which can be used as biosensing platforms. Many nanopatterning tech- niques require electron-beam or photo-lithography in vacuum environments [6], using a polymer mask [7,8], or stamping approaches [9], and cannot produce patterns on the nanoscale. The atomic force microscopy (AFM)-based nanopatterning technique simply involves using an atomic force microscope, where the AFM probe is used as a sharp stylus to scratch the thiols from the surface. The force applied by the AFM probe can easily disturb the sulfur bond between the thiols and metal surface. This approach has been demonstrated by producing simple defects in thiol monolayers [10]. This opened the development of a new nanografting method for patterning SAMs with nan- ometer precision [10]. In this study, we used the new nanografting method to produce a pattern by mechanically substituting one thiol with another using an AFM probe in the solution of the second thiol. The scratched squares © 2011 Moores et al; licensee Springer. This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited. Moores et al. Nanoscale Research Letters 2011, 6:185 http://www.nanoscalereslett.com/content/6/1/185 Page 2 of 5 produced by the AFM probe were immediately filled by the second thiol present in solution due to the higher che- mical concentration. This procedure makes it possible to create an alternating charge pattern composed of two dif- ferent thiols. We used the AFM to characterize surface morphology and Kelvin probe force microscopy (KPFM) to characterize the surface potential of the produced pattern. Results and discussion Nanografting Uniform surface coverage with one thiol was created by incubating a gold surface for 24-72 h in this thiol solu- tion. After incubation, the sample was exposed to a sec- ond thiol solution in an AFM liquid cell. The AFM probe, with a spring constant of 5 N/m, was inserted into the liquid cell and was used to scratch squares of defined dimension, varying from 10 by 10 nm to 10 by 10μ m in contact mode. The force applied was just high enough to remove thiol molecules from the surface (approximately between 10 and 50 nN), thus leaving the gold unda- maged. We have performed scratching at high speeds (10 lines/s) in order to reduce thermal drift and decrease the time required to create a pattern. The number of lines per square depends on the tip geometry, but we found for 10 μm2 squares at least 512 lines were required. The scratched squares produced by the AFM probe were immediately filled by thiol 2 present in solution due to the higher concentration of this thiol. The second thiol must have a higher affinity for the metallic surface to replace the first thiol removed from the surface by AFM probe. This process makes it possible to create an alter- nating charge pattern, composed of two different thiols. AFM and KPFM were used to characterize the pattern in terms of topography and surface potential. We first incubated a solution of CH3-terminated thiol molecules on gold-coated glass for 24 h. Figure 1a shows an AFM topography image of this thiol SAM in air. We applied a three-step nanografting method [11] to produce a pattern. First, AFM was used to image a previously formed monolayer (matrix SAM) in a solu- tion with another thiol (COOH-terminated thiol). Sec- ond, the tip was positioned into a selected spot to start a programmed scratching of defined areas. The scratch- ing was performed with a higher load than the threshold for thiol 1 (CH3-terminated) displacement [12]. During the scratching, the AFM probe removed matrix thiol 1 and produced bare gold squares exposed to thiol 2 (COOH-terminated) solution (nanoshaving) [13]. Surface potential of thiol SAM pattern Figure 2a shows AFM topography image of the two- thiols pattern, created by substituting thiol 1 (CH3- terminated thiol) with thiol 2 (COOH-terminated thiol). Figure 1 AFM topography of CH3 thiol SAM and pattern. AFM topography of (a) a uniformly covered CH3 thiol surface, and (b) a nanopattern shaved into a thiol surface exposing gold surface. Topography does not show much contrast as the two thiols do not differ significantly in height. The cross- section plot for topography image shows flat profile, with exception of few impurities. Figure 2b shows a sur- face potential map, obtained with KPFM and reveals a pronounced difference (20 mV) in surface potential on the border of two thiols (cross-section plot, Figure 2d). Moores et al. Nanoscale Research Letters 2011, 6:185 http://www.nanoscalereslett.com/content/6/1/185 Page 3 of 5 Figure 2 AFM and KPFM of CH3/COOH thiol pattern. Nanopattern (a) topography and (b) KPFM produced using CH3 and COOH thiols. (c,d) show cross-sections of the topography and KPFM, respectively. Conclusions In summary, we showed that thiol SAM pattern with chemical functionality and desired surface potential dif- ferences can be created using AFM-based nanografting method. In addition, we demonstrated for the first time that small differences in surface potential maps asso- ciated with organic thiol patterns can be resolved by KPFM in mV range. Such patterns with controlled dif- ferences in surface potential can be useful in nano- and bio-technology applications and to study interactions of Moores et al. Nanoscale Research Letters 2011, 6:185 http://www.nanoscalereslett.com/content/6/1/185 Page 4 of 5 charged species, such as nanoparticles and macromole- cular ions with non-uniformly charges surfaces. Methods Chemicals and sample preparation Decanethiol, cysteamine hydrochloride, 3-mercaptopro- pionic acid, and HPLC grade ethanol were purchased from Sigma-Aldrich Chemical Co. (St. Louis, MO, USA). These chemicals produce CH3, NH2, and COOH- terminated surfaces, respectively. All chemicals were used as received with no further purification. Thiols were dissolved in ethanol at 5 mM concentration. Thiol SAM preparation Gold-coated mica slides were purchased from Agilent Technologies, Inc. (Santa Clara, CA, USA). Before use, these gold surfaces were glued to clean glass cover slips using Epo-Tek 377 glue from Epo-Tek, Inc. (Billerica, MA, USA), which was cured at 150°C for 1 h [14]. The mica slide was removed from the "sandwich" substrate, leaving the glass with attached gold thin film, revealing the atomically flat gold side. The exposed gold was imaged to confirm atomically flat topography. The gold surfaces were then incubated in an appropriate 5 mM thiol solution in ethanol for 24-72 h to obtain uniform SAM surface coverage. Atomic force microscopy AFM uses a sharp probe over a sample surface and allows for imaging the topographical features at the nanoscale. Two common modes of operation are contact mode and intermittent contact mode. Thiol-modified surfaces were imaged in intermittent contact mode with a JPK Nanowi- zard II atomic force microscope. In intermittent contact mode, the tip is oscillated at the resonant frequency of the cantilever, and a feedback loop maintains constant amplitude over the entire image to insure the gentle ima- ging conditions. The probes used were Nanoworld NCH tips with a resonant frequency of approximately 338 kHz and 42 N/m spring constant. In contact mode of imaging, the tip usually lightly touches the surface and is moved up and down with the topographical features of the sam- ple. With increased force the probe can interact strongly with the surfaces and remove soft matter from the sur- face. This approach was used for nanografting. Auto- mated patterning was achieved by programming the JPK AFM imaging software to scratch a square of defined size and then move to the next defined location. Alternating this process produces the pattern with the size of few nm2 to few μm2. Kelvin probe force microscopy KPFM is an extension of AFM that provides the ability to map the surface potential in addition to imaging sample topography [15-17]. KPFM measures the surface potential by eliminating the electrostatic interactions between the tip and sample by applying a DC bias. This DC bias is tuned by a feedback loop that monitors mechanical oscillations induced in the tip due to an AC voltage (1 V) applied to the tip or sample. KPFM images were recorded using lift mode (also known as hover mode) operation. In lift mode KPFM, the topography of the sample is measured during the trace scan without an applied potential. During the retrace of the same line, the tip follows the topography measured during the trace pass but offset 50 nm above the surface, and an AC and DC voltage is applied between the tip and sam- ple to nullify the electrostatic interactions. Increasing the tip-sample separation by 50 nm eliminates the possi- bility of cross talk between the topography and surface potential measurements. Surface potential images were recorded in air using Nanoworld NCH cantilevers with a JPK Nanowizard II AFM in a hover mode KPFM. The gold substrates were grounded to eliminate sample charging. Abbreviations AFM: atomic force microscopy; KPFM: Kelvin probe force microscopy; thiols: SAMs (self-assembled monolayers). Acknowledgements The authors acknowledge technical support from JPK Instruments, Germany, and Agilent Technologies, USA. The authors acknowledge financial support from Natural Science and Engineering Council of Canada (NSERC), Canadian Foundation of Innovation (CFI), Ontario Research Fund (ORF), as well as Waterloo Institute for Nanotechnology (WIN) Graduate Scholarship Award to B. Moores. Author details 1Department of Physics and Astronomy, University of Waterloo, 200 University Avenue West, Waterloo, ON N2L 3G1, Canada. 2Department of Biology, University of Waterloo, 200 University Avenue West, Waterloo, ON N2L 3G1, Canada. 3Agilent Technologies, 4330 W. Chandler Blvd. Chandler, AZ 85226, USA. Authors' contributions BM and JS carried out the thiol SAM preparation and nanografting experiments, participated in the manuscript draft preparation. BM carried out KPFM imaging. SX participated in nanografting experiments and participated in the manuscript draft preparation. ZL conceived of the study, participated in its design and coordination and finished the final draft of the manuscript. All authors read and approved the final manuscript. Competing interests The authors declare that they have no competing interests. Received: 12 November 2010 Accepted: 1 March 2011 Published: 1 March 2011 References 1. Bigelow WC, Pickett DL, Zisman WA: Oleophobic monolayers. 1. Films adsorbed from solution in non-polar liquids. J Colloid Sci 1946, 1:513-538. Chechik V, Stirling CJM: Gold-thiol self-assembled monolayers. In The Chemistry of organic derivatives of gold and silver. Volume chapter 15. Edited by: Patai S, Rappoport Z. Hoboken NJ: John Wiley 1999:561-640. Ulman A: An Introduction to Ultrathin Organic Films: From Langmuir-Blodgett to Self-Assembly New York: Academic Press; 1991. 2. 3. Moores et al. Nanoscale Research Letters 2011, 6:185 http://www.nanoscalereslett.com/content/6/1/185 Page 5 of 5 4. 5. 6. 7. 8. 9. 10. Strother T, Hamers RJ, Smith LM: Covalent attachment of oligodeoxyribonucleotides to amine-modified Si (001) surfaces. Nucleic Acids Res 2000, 28:3535-3541. Boon EM, Salas JE, Barton JK: An electrical probe of protein-DNA interactions on DNA-modified surfaces. Nat Biotechnol 2002, 20:282-286. Baralia GG, Pallandre A, Nysten B, Jonas AM: Nanopatterned self- assembled monolayers. Nanotechnology 2006, 17:1160-1165. Chen W, Ahmed H: Fabrication of 5-7 nm wide etched lines in silicon using 100 keV electron-beam lithography and PMMA resist. Appl Phys Lett 1993, 62:1499-1501. Vieu C, Carcenac F, Pépin A, Chen Y, Mejias M, Lebib A, Manin-Ferlazzo L, Couraud L, Launois H: Electron-beam lithography: resolution limits and applications. Appl Surf Sci 2000, 164:111-117. Kumar A, Abbott NL, Kim E, Biebuyck HA, Whitesides GM: Patterned self- assembled monolayers and meso-scale phenomena. Acc Chem Res 1995, 28:219-226. Liu GY, Xu S, Qian Y: Nanofabrication of self-assembled monolayers using scanning probe lithography. Acc Chem Res 2000, 33:457-466. 11. Xu S, Liu GY: Nanometer-scale fabrication by simultaneous nanoshaving 12. and molecular self-assembly. Langmuir 1997, 13:127-129. Liu GY, Salmeron MB: Reversible displacement of chemisorbed n-alkane thiol molecules on Au(111) surface: an atomic force microscopy study. Langmuir 1994, 10:367-370. 13. Xu S, Laibinis PE, Liu GY: Accelerating the kinetics of thiol self-assembly on gold-a spatial confinement effect. Am Chem Soc 1998, 120:9356-9361. 14. Wagner P, Hegner M, Guntherodt HJ, Semenza G: Formation and in situ modification of monolayers on template-stripped gold surfaces. Langmuir 1995, 11:3867-3875. 15. Nonnenmacher M, O'Boyle MP, Wickramasinghe HK: Kelvin probe force microscopy. Appl Phys Lett 1991, 58:2921-2923. 16. Zerweck U, Loppacher C, Otto T, Grafstrom S, Eng LM: Accuracy and resolution limits of Kelvin probe force microscopy. Phys Rev B 2005, 71:125424-12543. 17. Moores B, Hane F, Eng L, Leonenko Z: Kelvin probe force microscopy in application to biomolecular films: frequency modulation, amplitude modulation, and lift mode. Ultramicroscopy 2010, 110:708-711. doi:10.1186/1556-276X-6-185 Cite this article as: Moores et al.: AFM-assisted fabrication of thiol SAM pattern with alternating quantified surface potential. Nanoscale Research Letters 2011 6:185. Submit your manuscript to a journal and benefi t from: 7 Convenient online submission 7 Rigorous peer review 7 Immediate publication on acceptance 7 Open access: articles freely available online 7 High visibility within the fi eld 7 Retaining the copyright to your article Submit your next manuscript at 7 springeropen.com
1309.0426
1
1309
2013-09-02T14:45:28
A molecular dynamics simulation of DNA damage induction by ionizing radiation
[ "physics.bio-ph", "cond-mat.dis-nn", "cond-mat.mes-hall", "physics.chem-ph", "physics.comp-ph" ]
We present a multi-scale simulation of early stage of DNA damages by the indirect action of hydroxyl ($^\bullet$OH) free radicals generated by electrons and protons. The computational method comprises of interfacing the Geant4-DNA Monte Carlo with the ReaxFF molecular dynamics software. A clustering method was employed to map the coordinates of $^\bullet$OH-radicals extracted from the ionization track-structures onto nano-meter simulation voxels filled with DNA and water molecules. The molecular dynamics simulation provides the time evolution and chemical reactions in individual simulation voxels as well as the energy-landscape accounted for the DNA-$^\bullet$OH chemical reaction that is essential for the first principle enumeration of hydrogen abstractions, chemical bond breaks, and DNA-lesions induced by collection of ions in clusters less than the critical dimension which is approximately 2-3 \AA. We show that the formation of broken bonds leads to DNA base and backbone damages that collectively propagate to DNA single and double strand breaks. For illustration of the methodology, we focused on particles with initial energy of 1 MeV. Our studies reveal a qualitative difference in DNA damage induced by low energy electrons and protons. Electrons mainly generate small pockets of $^\bullet$OH-radicals, randomly dispersed in the cell volume. In contrast, protons generate larger clusters along a straight-line parallel to the direction of the particle. The ratio of the total DNA double strand breaks induced by a single proton and electron track is determined to be $\approx$ 4 in the linear scaling limit. The tool developed in this work can be used in the future to investigate the relative biological effectiveness of light and heavy ions that are used in radiotherapy.
physics.bio-ph
physics
A molecular dynamics simulation of DNA damage induction by ionizing radiation Ramin M. Abolfath, David J. Carlson, Zhe J. Chen, Ravinder Nath Department of Therapeutic Radiology, Yale University School of Medicine, New Haven, CT, 06520-8040 (Dated: October 30, 2018) Purpose: We present a multi-scale simulation of early stage of DNA damages by the indirect action of hydroxyl (•OH) free radicals generated by electrons and protons. Methods: The computational method comprises of interfacing the Geant4-DNA Monte Carlo with the ReaxFF molecular dynamics software. A clustering method was employed to map the coor- dinates of •OH-radicals extracted from the ionization track-structures onto nano-meter simulation voxels filled with DNA and water molecules. Results: The molecular dynamics simulation provides the time evolution and chemical reactions in individual simulation voxels as well as the energy-landscape accounted for the DNA-•OH chemical reaction that is essential for the first principle enumeration of hydrogen abstractions, chemical bond breaks, and DNA-lesions induced by collection of ions in clusters less than the critical dimension which is approximately 2-3 A. We show that the formation of broken bonds leads to DNA base and backbone damages that collectively propagate to DNA single and double strand breaks. For illustration of the methodology, we focused on particles with initial energy of 1 MeV. Our stud- ies reveal a qualitative difference in DNA damage induced by low energy electrons and protons. Electrons mainly generate small pockets of •OH-radicals, randomly dispersed in the cell volume. In contrast, protons generate larger clusters along a straight-line parallel to the direction of the particle. The ratio of the total DNA double strand breaks induced by a single proton and electron track is determined to be ≈ 4 in the linear scaling limit. Conclusions: In summary, we have developed a multi-scale computational model based on first principles to study the interaction of ionizing radiation with DNA molecules. The main advantage of our hybrid Monte Carlo approach using Geant4-DNA and ReaxFF is the multi-scale simulation of the cascade of both physical and chemical events which result in the formation of biological damage. The tool developed in this work can be used in the future to investigate the relative biological effectiveness of light and heavy ions that are used in radiotherapy. PACS numbers: 82.50.-m, 87.50.-a, 87.53.-j I. INTRODUCTION Ionizing radiations (X/ γ-rays, α-particles, and heavy ions) induces damages to DNA-molecules such as single- strand breaks (SSB), double-strand breaks (DSB) and base damages (BD) via complex processes of direct ion- izations and/or indirect action by free-radicals with a ratio that is determined by the energy and the type of radiation source. In the indirect mechanism of radiation interactions, the radiation ionizes water molecules and creates neu- tral free-radicals and aqueous electrons (Ward 1988, Mozumder 1985, Oliveira 2012, Moiseenko 1998). The process involves the generation and diffusion of •OH rad- icals in biological and/or aqueous environments followed by chemical reactions that allow removal of hydrogen atoms from the DNA. This process is energetically fa- vorable for •OH radicals as it forms a water molecule and fills the electronic shell by neutralizing its magnetic moment. Great efforts have devoted to the statistical modeling of the damage sites based on Monte-Carlo (MC) sampling of the radiation track-structure and clustering of ioniza- tions, using empirical reaction rates and radiation scat- tering cross-sections (Aydogan 2008, El Naqa 2012, Moi- seenko 1998, Friedland 1999, Terrisol 1990, Wilson 1994, Nikjoo 1994, 1995, 1997, Zaider 1984, Semenenko 2005, Bernal 2009, Bernal 2011, Francis 2011, McNamara 2012, Kalantzis 2012, Michalik 1993, Brenner 1992, Lea 1946, Stewart 2011). Along this line we propose a first princi- ple quantum mechanical simulation using the molecular dynamics (MD) computational models to provide more information on the biological and chemical pathways of radiation-induced DNA damage. In the chemistry litera- ture, MD have shown great success for applications such as enzymatic reactions (Antoniou 2004, Karplus 1990, Karplus 2002, Schwartz 2009) and drug design (Durrant 2011). In nano-dosimetry studies, see e.g., (Moiseenko 1998, Bernal 2009, Bernal 2011), molecular structure of DNA and chromosomes has been used for scoring direct and indirect hits. In these models geometrical volumes are constructed to represent assemblies of atoms, e.g., the DNA-base. Alternatively atoms can be described by an effective and adjustable volume such that if energy de- position occurs in a volume occupied by assemblies of DNA-atoms (e.g., DNA bases) or water molecules in the hydration shell, a direct hit is recorded, and this partic- ular deposition is removed from further consideration. For DNA damage caused by •OH-radicals, an energy deposition has to occur within an adjustable diffusion length from the DNA. Coarse-grained reaction-diffusion rate models (Nikjoo 1994, 1995, 1997, Zaider 1984) have been developed to describe the process. In the atom- 3 1 0 2 p e S 2 ] h p - o i b . s c i s y h p [ 1 v 6 2 4 0 . 9 0 3 1 : v i X r a istic level, the calculation of the diffusion length requires combination of quantum mechanics (QM) and molecu- lar mechanics (MM). Preliminary results using QMMM (Abolfath 2012) show •OH-radicals close to 1 nm from the surface of DNA can reach the DNA within a ps time scale and perform subsequent chemical reaction. The average diffusion distance of a hydroxyl radical in a cellular milieu has been measured to be about 6 nm (Roots 1975). This is approximately three times the di- ameter of the DNA double helix, which implies that the chance an •OH to damage the DNA decreases rapidly with distances beyond about 6 nm. To draw an accurate conclusion on the diffusion length predicted by MD and for comparison with the experiments, a systematic cal- culation must be carried out along the previous work to complete the results. For the current problem of interest, we introduce an alternative computational approach that permits going beyond phenomenological models and volumetric ap- proaches such as those presented in Refs. (Aydogan 2008, El Naqa 2012, Moiseenko 1998, Friedland 1999, Terrisol 1990, Wilson 1994, Nikjoo 1994, 1995, 1997, Zaider 1984, Semenenko 2005, Bernal 2009, Bernal 2011, Francis 2011, McNamara 2012, Kalantzis 2012, Michalik 1993, Brenner 1992,L eaBook) as well as QMMM approaches such as Ref. (Abolfath 2012). II. MATERIALS AND METHODS: A selective sampling of the DNA damage caused by the ionization track structure is performed by partition- ing the cm-size MC volume into a three dimensional lat- tice of 1019 simulation voxels (SVs) with dimension 2.6 × 2.6 × 6 nm3. In each SV, we placed a fragment of DNA with the helical axis parallel to the beam direction and fill the empty space of the SV with water molecules. Geant4-DNA MC simulation is used to calculate the spa- tial distributions of ions in and among SVs and to char- acterize the dependence of DNA-damage pattern on the beam source and quality. To generate the coordinates of the ions on a nano- scale level, we employ the GEANT4-DNA extension of the GEANT4 Monte Carlo simulation toolkit, version 9.6.p01-64bits, (Incerti 2010a, 2010b). The GEANT4- DNA package allows the event-by-event simulation of the particle shower produced during the transport of elec- trons and protons in a continuous model of liquid water. In the current version of Geant4-DNA, the static struc- ture of the water molecules is taken from the scattering cross sections. However, the molecular representation of water included in the physics lacks the dynamical as- pects needed for studying the time-evolution of ioniza- tion. Moreover the DNA structure is introduced as a target volume inside a water phantom, pretending that the target volume is filled by DNA (Bernal 2011). In our approach, we assume that at the location of ionized water, an •OH free radical is created. By engaging MD 2 FIG. 1. (color online). A sample of ionization calculated by Geant4-DNA for a single-track of electron (a-b) and proton (c-d) in nano-meter scale. For illustration a segment of DNA parallel to the direction of central axis of the beam is added. Ions generated by the beam of electron are scattered in space. In comparison, ions generated by the beam of proton are in a straight line along the initial direction of proton. At the point of ionizations we introduce a diatomic •OH-radical with a random diatomic orientation. simulation, we calculate the dynamical trajectory and the energy-landscape of the DNA molecule in the pres- ence of water and •OH-radicals on-the-fly and hence the first-principle enumeration of DNA lesions driven by cas- cade of chemical reactions can be carried out. Here, and throughout this work, a DNA lesion refers to a single SSB or BD. The energy-landscape of the system is the total potential energy of DNA and •OH, calculated as a func- tion of reaction distance, e.g., the distance between DNA and the transferred hydrogen to •OH. This hydrogen transfers to •OH and form a water molecule. As a result of collective migration of hydrogen from DNA, cascade of chemical reactions such as carbonyl- and hydroxyl- hole formation in the sugar-moiety rings, nucleotide- nucleotide hydrogen bond disruption, nucleotide struc- tural damage and nucleotide-sugar-moiety bond breaking occur. (a)(b)33nmRadiation source: electron(c)(d)200 nmRadiation source: proton 3 FIG. 4. (color online). Histogram of ionizations distributed among nm-size SVs for a single electron and proton track (red tracks shown in Figs. 2 (a) and (b), respectively). The total number of SVs occupied by at least one ionization is 46,000 and 28,000 for an electron and proton, respectively. FIG. 2. (color online). Four ionization-tracks of 1 MeV electrons (a) and two ionization-tracks of 1 MeV protons (b) obtained from the Geant4-DNA Monte Carlo simulation in water with initial points at x = y = z = 0. Colors rep- resent tracks by a series of uncorrelated initial seeds in the Geant4 random number generator. The length is in micro- meter (mm) and the arrows show the initial direction of the beam. FIG. 5. (color online). Final structure of DNA in solution after 10 ps MD simulation at room temperature. For clarity in visualization, in (b) the water molecules are removed from the identical computational box shown in (a). Atoms color code: carbon, oxygen, nitrogen, phosphorus and hydrogen atoms are shown as cyan, red, blue, gold and white, respectively. We further investigate the dependence of DNA-•OH re- activity on the formation of hydrogen-peroxide com- pounds and the network of •OH··· •OH hydrogen bonded molecules. We demonstrate the stability of these chem- ical complexes and show that they prevent free radicals from reaching the DNA molecule. Such complexes tend to form for •OH-radicals in close proximity to each other, i.e., in sub-critical clusters with dimension less than a critical value with radius smaller than 2-3 A. For a pair of •OH radicals the radius of a sub-critical cluster is de- termined by the hydrogen-bond effective length as two FIG. 3. (color online). Track averaged radial density of ionizations as a function of distance from the source. Track averaging was performed over ten 1 MeV electrons and pro- tons. The ionization density generated by protons shows a peak at the proton range, resembling the Bragg peak. The total number of ionizations per track for both electrons and protons is ≈50,000. e-ݔሺߤ݉ሻݕሺߤ݉ሻݖሺߤ݉ሻ െ  "&െ"(a)െ&െ&pݔሺߤ݉ሻݕሺߤ݉ሻݖሺߤ݉ሻ  ##"&െ"(b)Ionizations per µmelectron single beam (1MeV)proton single beam (1MeV)10-4 10-3 10-2 10-1 100r(cm)10410310210110010-10200040006000Simulation voxel number051015202530Ionizations per SV1 MeV single e beam1 MeV single p beam(a)(b) 4 (color online). Average number of DNA broken FIG. 7. bonds, ¯Lα N, as a function of number of ionizations N per SV, calculated for a single electron (e) and proton (p) track (red tracks shown in Fig. 2 (a) and (b), respectively). Solid line shows the linear scaling limit, ¯Lα N = N where one ion gener- ates one DNA broken bond, neglecting sub-critical clusters. The deviation from linear scaling for electron is clearly more pronounced than for proton. 4GB RAM is sufficient to run ReaxFF-MD program em- bedded in LAMMPS. The program is fast and efficiently parallelized. The speed depends on the time steps and varies between few hours to few days. The consistency and accuracy of ReaxFF with local density functional calculation is checked systematically by employing ab- initio Car-Parrinello MD (Car 1985, Hutter 2008). In ReaxFF the atomic interactions are described by the re- active force field potential (van Duin 2001). ReaxFF is a general bond-order dependent potential that provides ac- curate descriptions of bond breaking and bond formation. Recent simulations on a number of hydrocarbon-oxygen systems (Bagri 2010), and organic molecules (Abolfath 2011, Monti 2012) showed that ReaxFF reliably provides energies, transition states, reaction pathways and reac- tivity trends in agreement with ab-initio calculations and experiments. In our method, the spatial distribution of the ionized water molecules and •OH-radicals are obtained by post- processing of the track structures calculated by Geant4- DNA. A series of statistically uncorrelated single particle tracks are generated and used for the sampling of the DNA lesions by changing the random number generator seeds. We specifically focus on single electron/proton beams with initial energy of 1 MeV as an illustration of the methodology. We obtain the coordinates and spatial distribution of water ionizations in a cm-size phantom for each track. For the statistical sampling of •OH in SVs, for each track we enumerate the number of ions per SV, N, and the number of SVs, KN , for a given N. The chemical interaction between hydroxyl radicals and DNA and the scoring of hydrogen abstractions, chemi- (color online). FIG. 6. (a) Initial distribution of water molecules and randomly generated small pockets of •OH- radicals surrounding the DNA (t=0). The dimension of the MD computational box (SV) is 2.6 × 2.6 × 6 nm3. (b) For visual clarity, water molecules and •OH-radicals are removed from the identical computational box shown in (a). (c) Final configuration of DNA at t=10 ps in water and in the pres- ence of •OH-radicals. (d) Same as (c) after deleting water and •OH molecules. The white dots in the background are abstracted hydreogens from DNA. The color code used to la- bel the atoms is cyan, red, blue, gold and white for carbon, oxygen, nitrogen, phosphorus and hydrogen, respectively. radicals can form •OH··· •OH complex (here ··· repre- sents hydrogen bond) or a covalent bond length if they form hydrogen peroxide, HO-OH. In ReaxFF-MD, for- mation of •OH··· •OH and HO-OH bonds are seen for a distance between two •OH radicals less than 2-3 A. Our approach for reactive MD consists of ReaxFF (van Duin 2001, Chenoweth 2008) implemented in Large- scale Atomic/Molecular Massively Parallel Simulator, LAMMPS, version Jan. 2011 (Plimpton 1995, 1997). In regard to the software performance in terms of comput- ing time and memory usage, a modest computer with (a)(b)(c)(d)0246810121416Number of ionization per SV024681012Average broken bonds 1 MeV e1 MeV p cal bond breaking, and DNA lesion formation is simu- lated by ReaxFF-MD. We fill the SVs with DNA and water molecules corresponding to the density equal to 1 g/cm3 using the computational tools available in GRO- MACS (Berendsen 1996, Lindahl 2001, Van Der Spoel 2005, Hess 2008, Abolfath 2012). GROMACS is a versa- tile package to perform classical molecular dynamics. It simulates the Newtonian equations of motion for systems with hundreds to millions of particles. An interface between the output of Geant4-DNA with the input of ReaxFF-MD (Abolfath 2011, van Duin 2001) was constructed to convert the coordinates of ionization events obtained in Geant4-DNA to generate the coor- dinates of diatomic structure of •OH-radicals used in ReaxFF-MD at the ionization points with random an- gular orientations with respect to the DNA-molecule, as shown in Fig. 1. A series of ReaxFF-MD were performed to calculate the average number of DNA lesions per SV per beam for a given N ionizations. A DSB consists of two or more SSB on opposite strands of the DNA that are spatially and temporally close (Ward 1988). As shown in Fig. 1, we construct a large fragment of a solvated 950-atom clas- sical Watson-Crick DNA-strand (Munteanu 1998), con- sisting of 15 base-pairs, with approximately 1,100 wa- ter molecules added using GROMACS (Berendsen 1996, Lindahl 2001, Van Der Spoel 2005, Hess 2008, Abolfath 2012) along the principal axis of the SV, the z -axis. The vertical dimension of the SV and the length of the DNA molecule are optimized to balance between the min- imum size of a DNA to fit one DSB per DNA-length and the largest fragment of DNA including the surrounding water molecules, allowable in the MD simulation. The diffusibility of •OH radicals must be taken into account to estimate a reasonable cut-off on the SV lateral dimen- sions. When the lateral size of SV increases, the number of water molecules surrounding the DNA increases and induces an unnecessary increase of the simulation time. Consequently, we chose the lateral dimension of the SV equal to 2.6 nm. To minimize the computational time, it is favorable to choose SVs as small as possible. However, by choosing very small SVs, we may lose layers of water molecules that are within the •OH diffusion length. Therefore, an optimal value of the lateral dimension of the SV is de- termined by optimizing the number of water molecules surrounding the DNA to incorporate properly the diffu- sion of •OH-radicals within ps time-scale. We use the constructed molecular structure as a replica for all SVs •OH-radicals are added in the Geant4-DNA volume. in the position of ionization coordinates obtained from Geant4-DNA with random orientation relative to the DNA. These simulations were performed using periodic boundary conditions in a canonical moles, pressure and temperature (NPT) ensemble, with a Nose-Hoover ther- mostat for temperature control and a time step of 0.25 fs and run MD for 50 ps, as described in Ref. (Abolfath 2011). Within this period of time, •OH radicals interact 5 with DNA to induce DNA lesions, i.e., SSB and BD, and subsequently DSB. Two different sets of MD simulation were performed: in the presence and absence of •OH-radicals. We follow the molecular dynamics of •OH radicals in the presence of DNA and water using molecular time-evolution built in the ReaxFF-MD. Therefore, all events in our approach are accounted for as indirect damage by the •OH free radical. The relevant time-scales for the sequence of sim- ulations, consistent with the experimental data (El Naqa 2012,Moiseenko 1998, Sies 1993) ranges from atto- to nano-seconds, relevant to MC and MD events, respec- tively (Mozumder 1985). From the output of the MD calculation, we count the number of broken DNA bonds resulting in DNA lesions and DSBs for each track. The above procedure is then repeated to obtain the final re- sults that are averaged over all simulated tracks. III. RESULTS: Fig. 2 shows samples of the ionization-tracks of 1 MeV electrons (a) and protons (b) obtained from the Geant4- DNA Monte Carlo simulation in a continuous model of water. The maximum beam energy in the current version of Geant4-DNA is 1 MeV. The arrows show the initial di- rection and origin of the particles. Each point represents a single ionization event. Although the total number of ionizations produced by a single 1 MeV electron and pro- ton track is approximately the same (about 50,000), the range of a 1 MeV proton is clearly shorter than the range of a 1 MeV electron due to the greater density of ioniza- tions along the proton track. Because the electron un- dergoes stronger lateral scattering by the medium com- pared to the proton, we observe a strong divergence of the ionization tracks of electrons in space compared to protons. Moreover, a large fraction of the ions produced by the electrons are in the negative direction relative to the direction of the incident electron, indicating that the number of back-scattering events for electrons is much greater than for protons. Fig. 3 illustrates the radial density of ionizations av- eraged over ten tracks and tallied into one micro-meter radial slices. The distance from the source, r, is given for the ionization tracks shown in Fig. 2. The average ionization range of 1 MeV electrons and protons is 3000 and 25 µm, respectively. The calculated ionization range of electron is consistent with the NIST-reported CSDA range of a 1 MeV electron of ≈ 4 mm. The latter in- cludes all excitations including ionization events. Proton ionization density shows a peak in the tail of the track, resembling the Bragg peak. The average depth of the range of the electrons occurs at a distance two orders of magnitude larger than the average proton range be- cause the mass of the electron is three orders of magni- tude smaller than the proton. The maximum number of ionizations per micro-meter length per track accounted for the proton beam is three orders of magnitude larger than that for the electron beam. The calculated total ionization energy deposited by the electrons and protons per track is 660 and 640 keV, respectively. Hence, we find the electron and proton stopping powers averaged over the entire track to be approximately 0.22 and 26.6 keV/µm, respectively, which is consistent with published values (Bernal 2011). Fig. 4 shows a typical distribution of ionizations among 2.6 × 2.6 × 6 nm3 MD-SVs calculated for a single 1MeV electron and proton track (red tracks shown in Fig. 2 (a) and (b), respectively). There are 1019 SVs that cover the three-dimensional structure of the cm-size MC volume. The number of ions per SV per track, N, is extracted from post-processing of the output of Geant4-DNA. N is an integer number that ranges from zero to Nmax. Be- cause a single particle track deposits ions in only a small fraction of the SVs, the distribution is highly sparse, i.e., there is a large number of SVs with N=0. As we de- fined KN is the number of the SVs that are occupied by N ions. This quantity partitions the SVs based on the number of ionizations. As pointed out above, the total number of ionizations deposited by the beam of electron is approximately the same as the beam of proton. For clarity, the number of ionizations shown in the x -axis of Fig. 4 is truncated at 6,000. The areas under the curves are the total number of ionizations. Figs. 5 and 6 show the structure of the DNA and wa- ter molecules in the ReaxFF-MD computational box. In SVs with zero ionization (N=0 ), we do not expect any DNA-damage. However, validation of the ReaxFF-MD results can be demonstrated by performing molecular dy- namics for this situation. The simulation illustrated in Fig. 5 describes equilibration of the solvated DNA-strand at T=300K for 50 ps. We found that during this time- frame the DNA retained its overall helical configuration, indicating that ReaxFF retains the overall structural in- tegrity of the DNA over such time-frames and that re- active events observed during exposure to •OH radicals, as shown in Figs. 6 (a-d), can indeed be associated with the radical reactivity. Fig. 6 (b) illustrates the initial DNA structure used in MD simulation. During the simulation time, chemical re- actions between •OH-radicals and DNA lead to DNA hy- drogen abstraction. This can be observed as white dots in the background of Fig. 6 (d). In the final state, i.e., Figs. 6 (c) and 6 (d), a pronounced separation between bases and backbone is visible. Strong distortion in base stack- ing with large holes are clearly seen. Large separation between bases and the backbone as well as the stretched backbone illustrating a collective base-damage, and large number of single-strand breaks, are visible. As a result of chemical reactions, the charge distribution on the DNA is dramatically altered and the helical-structure becomes unstable. Consequently, a series of bonds are broken. In Fig. 6 (d) we observe number of missing links between atoms, accounted for broken covalent bonds. Hence, comparing Figs. 6 (b) and (d), we can clearly count the number of missing covalent bonds accounted for DNA le- 6 sions, as each broken covalent bond corresponds to one DNA lesion. Moreover, we identify chemical pathways for carbonyl- and hydroxyl-hole formation in the sugar- moiety rings, nucleotide-nucleotide hydrogen bond dis- ruption, and nucleotide structural damage throughout a sequence of intermediate events such as DNA-backbone and -base hydrogen abstraction by •OH-radicals followed by nucleotide-sugar-moiety bond breaking. Figure 7 shows the average number of DNA broken bonds, ¯Lα N per SV, as a function of the number of ion- izations N per SV and the particle-type, α={e, p}, cal- culated by ReaxFF-MD for the red-tracks in Fig. 2. In the limit of small N 's, and for both electron and pro- N ≈ N. However, a deviation from linear scaling ton, ¯Lα emerges for N = 3 due to the formation of sub-critical clusters of ions. The results obtained directly from MD simulation show that the ions localized in these clusters can form only a single broken bond in DNA, hence they suppress biologically effectiveness of the beam as argued by Lea (Lea 1946). The number of SVs, KN , for a given N can be calculated from Fig. 4. For all tracks of electrons considered in this study, we find that 95% of the ions contribute to single ionization per SV, i.e., the number of SVs that are occupied only with a single ion, N=1, forms 95% of the population of the ions. This ratio drops to 80% for the proton track. Because the DNA broken bond induced by a single ion is either on one side of DNA- backbone or on DNA-base, 95% and 80% of the ions in the track of a single electron and proton, respectively, do not participate in DSB formation. Equivalently 5% and 20% of the electron and proton track, respectively, populate SVs with at least N = 2 and hence they can potentially participate in DSB formation. For N=2, the number of SVs containing sub-critical clus- N=2 ≈ 2. Considering only con- ters is small, hence ¯Lα 2 ≈ 1.02 where tribution of the N=2 -ions yields Kp Kp 2) is the population of the SVs with two ions gen- erated from a single particle track of proton (electron). The total number of DSBs generated by a beam of ioniz- ing radiation for a single track structure was calculated by convoluting the average number of lesions induced by a given number of ions, N, per SV, ¯Lα N, and their popu- lation in SVs, Kα N: 2 (Ke 2/Ke N α f α N max(cid:88) (cid:80)Kα N =0 DSBα = ¯Lα N Kα N. (1) N i=1 Lα max N =2 f α N N = 1 Kα N i where Lα i (cid:80)Kα (cid:80)N α Here f α N is a fraction of broken bonds propagating to DSBs and ¯Lα is the num- ber of broken-bonds in the i th-SV, hence DSBα = N = 1/2 if each pair of broken bond leads to a single DSB. The number of DNA lesions in a SV, Lα i , depends on both the number of •OH-radicals, N, and their spatial distribution. Lα i fluctuates among Kα N SVs because of the variation in the relative position of DNA and •OH radicals that influence . Note that f α N i=1 Lα i the diffusion length and •OH-radicals among each other with the possibility of the occurrence of the sub-critical clusters. Calculation of Eq.(1) requires a series of MD simulations. For example, for a single 1 MeV electron and proton track (red track shown in Figs. 2 (a) and (b), respectively), it requires KN = 1, 2, 3, 7, 7, . . . and KN = 29, 58, 84, 129, 147, . . . ReaxFF calculation, both correspond- ing to N = 16, 13, 12, 11, 10, . . . (see Fig. 4). Clearly KN increases as N approaches to zero, hence performing ReaxFF-MD simulation for all KN is a computationally expensive problem. To speed up the calculation, we per- form ReaxFF-MD for SVs, selected randomly from the beginning, middle, and the tail of the ionization tracks. We then calculate Lα N by averaging over the SVs for given tracks following by a second averaging over the simulated tracks. The average DSB per track, hence can be calculated by N and Kα max(cid:88) N α (cid:10)f α N (cid:11) , (cid:104)DSBα(cid:105) = α α N =0 (2) ¯Lα N Kα N where (cid:104)...(cid:105) denotes track-averaging. The dependence of the total number of ionizations on the number of tracks is approximately linear, assum- ing that the ionization-tracks are statistically uncorre- ∝ Ntracksfor small lated. Hence, the total DSBs, DSBtot Ntracks. Here Ntracks is the number of tracks. How- ever, the dependence of the number of DSBs on the number of tracks for both electrons and protons deviates from linear scaling owing to the saturation of the ions and increase in the population of the sub-critical clus- ters. With increase of Ntracks, the sub-linear behavior of DSBtot emerges, hence it asymptotically saturates to a maximum value. The onset of DSB saturation for proton appears in smaller Ntracks compare to electron as can be anticipated from Fig. 2. In the linear scaling limit, where a system of multi-tracks reduces to a single-track prob- lem, we find that the ratio of total DSB yields induced by a single electron and proton track with initial energy of 1 MeV, DSBp/DSBe, is 4. Note that in this limit (cid:104)DSBα(cid:105) ≈ DSBα where in the left hand side (cid:104)DSBα(cid:105) de- notes DSBs averaged over multi-tracks and DSBα in the right had side denotes the average of DSBs calculated from a single track. The upper limit of DSBp/DSBe can be calculated by ne- glecting the contribution of the sub-critical clusters. In this case we assume that a pair of ions generate one DSB, f α N = 1/2, and the number of DNA lesions scale linearly with number of ionization, ¯Lα N = N . In Fig. 7, the solid line shows such limit. It follows max(cid:88) N α N =0 DSBα = mod( N 2 ) Kα N, (3) where mod(N/2) is the integer part of N/2. For N=1 the number of DSBs is identical to zero. For N=2, Eq. 7 (3) predicts DSBα = Kα 2 since mod(N/2)=1. Clearly because Eq.(3) neglect the sub-linear behavior of DSB due to the emerging of the sub-critical clusters, it over- estimates the contribution of the ion-pairs thus it leads to a higher values for DSBp/DSBe than predicted by MD simulation of Eq.(1). For the case of 1 MeV beams, Eq.(3) predicts DSBp/DSBe = 4.4 that is 10% greater than the one calculated by Eq.(1). IV. DISCUSSION AND CONCLUSION: The DNA double strand break is the most biologi- cally relevant damage induced by ionizing radiation. To this end, we built an interface to export the output of Geant4-DNA to the ReaxFF-MD environment. The steps included (a) dividing a cm-size Geant4-DNA com- putational box into ≈1019 nm-size MD-SVs (b) construc- tion of a molecular structure of SVs under irradiation by adding water molecules as well as diatomic structure of •OH-radicals at the ionization points with random angu- lar orientations with respect to 15-bp Watson-Crick DNA molecules (c) calculation of inter- and intra-SV distribu- tion and numbers of •OH-radicals and (d) performing MD to enumerate DNA lesions. The range of energy used in this study is relevant to the low energy portion of a proton beam that results in a Bragg peak and is used in radiotherapy to deliver highly conformal dose distributions to tumors. The maximum particle energy was 1 MeV for both protons and electrons and the ionization events were collected from both single- track and multi-track beams. This maximum energy cor- responds to the maximum energy available in the current version of Geant4-DNA for electron. However, our ap- proach can be extended in the future to more clinically- relevant electrons and protons with energy greater than 1 MeV. In MD simulation, we assumed that randomly distributed clusters of diatomic •OH radicals are the main source of hydrogen abstraction. We demonstrated formation of carbonyl- and hydroxyl-groups in the sugar-moiety cre- ate holes that grow up slowly between DNA-bases and DNA-backbone and the damage collectively propagates to DNA single and double strand breaks. The time evo- lution of the entire system reveals chemical pathways for carbonyl- and hydroxyl-hole formation in the sugar- moiety rings as well as base damage induced by •OH- radicals. The clustering method of ionization in the simulation of track-structure can be found in the literature, including in Refs. (Michalik 1993, Brenner 1992). The method is based on the search for energy-deposition in spherical- shaped clusters containing certain number of ionizations. For example the K-means clustering of ionization intro- duced in Ref. (Michalik 1993) focuses on the partitioning of set of N ionizations into K roughly spherical clusters (Michalik 1993). By fitting to empirical DSB yield, one may correlate the frequencies of occurrence of clusters with a given size and given number of ionizations to rel- ative biological effects. Throughout such analysis it has been argued that locally multiply damaged sites are prob- ably caused by energy depositions producing at least 2 to 5 ionizations localized, respectively, in sites of diame- ters of 1 to 4 nm (El Naqa 2012, Michalik 1993, Brenner 1992). A central assumption in these other approaches is the mathematical mapping between the number of ioniza- tions localized in a cluster less than a critical size (sub- critical clusters) to a single site of damage in the track structure. In the track-structure simulation, if such a cluster is found, only one ionization inside the cluster is effectively considered to have caused the damage. All other ionizations inside that cluster are removed from further consideration, on the grounds that the site is already damaged. Mathematically this is equivalent of mapping all ionizations localized in a sub-critical clus- ter to a single damage, a homomorphism between the space of ionizations in the track-structure and the space of DNA damaged-sites. This is the phenomenon of satu- ration, discussed by Lea (Lea 1946), i.e., if multiple en- ergy deposition sites occur very close together (e.g., at high LET) they become less biologically effective. In the approach presented in this work, we relaxed the above as- sumption because the occurrence of a single DNA lesion by ions in sub-critical clusters can be explicitly modeled in MD simulation. Furthermore, two beams of electrons and protons can be differentiated by the distribution of ions in the SVs and the type of DNA lesions they produce. Hence, we at- tempt to correlate the DNA lesions and the beam source by performing a first-principle calculation via hybridizing MC and MD simulations. Another effect not considered in previous works is the reaction of radical products with each other which could reduce the hydroxyl radical yield (Brenner 1992). In our MD simulation, we can easily trace the dynamics of •OH radicals. We demonstrated the interaction among •OH radicals within a cluster that leads to the formation of hydrogen-peroxides and net- work of OH··· OH hydrogen bonded molecules as well as recombination of •OH and H to form stable water that significantly lowers the reactivity of DNA-•OH (Abolfath 2011). Our study reveals a qualitative difference in the DNA damage induced by low energy electrons and protons. Electrons mainly generate small pockets of •OH-radicals, randomly dispersed in the SV volume. In contrast, pro- tons generate larger clusters along a straight-line parallel to the initial direction of the particle. A quantitative comparison carried out on the DNA lesions in the limit of small track numbers shows that the total number of DSBs induced by a 1 MeV proton track is approximately four times greater than a 1 MeV electron track. How- ever, our preliminary observations indicate that this ra- tio may decrease because of the non-linearity on the de- pendence of the DSBs on the number of tracks, owing to the formation of the sub-critical clusters that occur 8 with higher frequency as the number of tracks increases. From the track structures shown in Figs.2 (a) and (b), one may expect that the onset of DSB saturation for pro- tons is reached sooner than for electrons. Whether or not (cid:104)DSBp(cid:105) /(cid:104)DSBe(cid:105)approaches unity, the clinically reported RBE values (Paganetti 2003), requires further investiga- tion. We note that in the current study we did not investi- gate the direct action, i.e., the direct ionization of DNA sugar-phosphate backbone, and the dissociation of water molecules into chemical species due to high energy events involving solvated electrons as discussed in (Karamitros 2011). We also considered only simulating hydroxyl free radicals and did not take into account the contribution of DNA damage induced by other free radical species. Al- though the dissociation of water molecules into charged ions such as H3O+ and OH− are embedded in our sim- ulation but as the current version of ReaxFF is not a suitable computational platform for processes involving solvated electrons, a full ab-initio method such as Car- Parinello MD (Car 1985, Hutter 2008) and QM-MM methods (Abolfath 2012) are also necessary. We post- pone our systematic study on these issues to forthcom- ing publications, however, spite of such limitation, we explored a number of interesting results using ReaxFF. In summary, we have developed a first-principle computa- tional model to study the interaction of ionizing radiation with DNA molecules at the microscopic level. As a gen- eral consensus, ionizing radiation changes the chemical environment of the DNA that subsequently leads to the chemical reactions in atomic level with manifestation in macroscopic scale, e.g., the DNA damage in the cellular level is the result of collective chemical reactions among chemical agents induced by ionizing radiation and DNA. As with any other collective phenomenon in physics, this problem can be studied in various scales, either at the macroscopic scale or in the atomic scale. In this study, we have taken the initial steps, as a proof-of-principle, to address the DNA-damage at the atomic level. Future works must be done to scale up the microscopic picture to the cellular level that experimental data are accessible. The advantage of the hybrid MC modeling of ionizing radiation with Geant4-DNA and MD of DNA-molecules using ReaxFF is based on the feasibility in performing a multi-scale simulation of the cascade of events in physical and chemical pathways with biological endpoints of the irradiated cells, a complex system that comprises DNA molecules in reactive environment. V. ACKNOWLEDGEMENT RMA thanks Nicole Ackerman, Satya Kumar, Moham- mad Kohamdel, Siv Sivaloganathan and Adri van Duin for useful discussion. VI. REFERENCES Abolfath R M, van Duin A C T, Brabec T 2011 Re- active Molecular Dynamics Study on the First Steps of DNA Damage by Free Hydroxyl Radicals J. Phys. Chem. A 115, 11045. Animations and movies are available in: http://qmsimulator.wordpress.com/ Abolfath R M, Biswas P K, Rajnarayanam R, Brabec T, Kodym R, Papiez L 2012 Multiscale QM/MM Molec- ular Dynamics Study on the First Steps of Guanine- Damage by Free Hydroxyl Radicals in Solution, J. Phys. Chem. A 116, 3940. Aydogan B, Bolch W E, Swarts S G, Turner J E, and Marshall D T 2008 Monte carlo simulations of site- specific radical attack to DNA bases, Radiat. Res. 169, 223-231. Antoniou D, Abolfath R, and Schwartz S D, 2004 Tran- sition Path Sampling Study of Classical Rate Promoting Vibrations, J. Chem. Phys. 121, 6442. Bagri A, Mattevi C, Acik M, Chabal Y J, Chhowalla M, and Shenoy V B 2010, Nature Chem., 2, 581. Berendsen H J C, van der Spoel D, van Drunen R 1996 Comp. Phys. Comm. 91, 43-56. Brenner D J, Ward J F 1992 Constraints on energy deposition and target size of multiply-damaged sites as- sociated with DNA double strand breaks, Int. J. Radiat. Biol. 61, 737-748. Bernal M A and Liendob J A 2009 An investigation on the capabilities of the PENELOPE MC code in nan- odosimetry, Med. Phys. 36, 620-625. Bernal M A, de Almeida C E, Sampaio C, Incerti S, Champion C, Nieminen P 2011 The invariance of the total direct DNA strand break yield", Med. Phys. 38, 4147- 4153. Car R, Parrinello M 1985 Phys. Rev. Lett., 55, 2471. Carlos S, Oliveira B, and Oliveira-Brett A M 2012 In Situ DNA Oxidative Damage by Electrochemically Gen- erated Hydroxyl Free Radicals on a Boron-Doped Dia- mond Electrode, Langmuir 28, 4896-4901. Chenoweth K, van Duin A C T, and Goddard W A 2008, ReaxFF reactive force field for molecular dynamics simulations of hydrocarbon oxidation: Journal of Physi- cal Chemistry A, 112, 1040-1053. El Naqa I, Pater P and Seuntjens J 2012 Monte Carlo role in radiobiological modelling of radiotherapy out- comes, Phys. Med. Biol. 57, 75. Francis Z, Incerti S, Capra R, Mascialino B, Montarou G, Stepan V, Villagrasa C 2011 Molecular scale track structure simulations in liquid water using the Geant4- DNA Monte-Carlo processes, Applied Radiation and Iso- topes 69, 220-226. Friedland W, Jacob P, Paretzke H G, Merzagora M and Ottolenghi A 1999 Simulation of DNA fragment distri- butions after irradiation with photons, Radiat. Environ. Biophys. 38, 39. Hutter J, Ballone P, Bernasconi M, Focher P, Fois E, Goedecker S, Parrinello M, Tuckerman M E, 2008 CPMD 9 fuer Festkoerperforschung, code, version 3.13, MPI Stuttgart IBM Zurich Research Laboratory. Hess B, van der Spoel D, Lindahl E 2008 J. Chem. The- ory Comput., 4, 435-447. Incerti S, Baldacchino G, Bernal M, Capra R, Cham- pion C, Francis Z, Gueye P, Mantero A, Mascialino B, Moretto P, Nieminen P, Villagrasa C, and Zacharatou C 2010a, The GEANT4-DNA project, Int. J. Model. Simul. Sci. Comput. 1, 157178. Incerti S, Ivanchenko A, Karamitros M, Mantero A, Moretto P, Tran H N, Mascialino B, Champion C, Ivanchenko V N, Bernal M A, Francis Z, Villagrasa C, Baldacchino G, Gueye P, Capra R, Nieminen P, and Zacharatou C 2010b, Comparison of GEANT4 very low energy cross section models with experimental data in water, Med. Phys. 37, 46924708. Karamitros M, Mantero A, Incerti S, Friedland W, Baldacchino G, Barberet P, Bernal M, Capra R, Cham- pion C, EL Bitar Z, Francis Z, Gueye P, Ivanchenko A, Ivanchenko V, Kurashige H, Mascialino B, Moretto P, Nieminen P, Santin G, Seznec H, Tran H N, Villagrasa C and Zacharatou C 2011, Modeling radiation chemistry in the Geant4 toolkit, Prog. Nucl. Sci. Tech. 2, 503-508. Lea D E 1946 Actions of Radiations on Living Cells, (Cambridge University Press, Cambridge). Lindahl E, Hess B, van der Spoel D 2001 J. Mol. Model. 7, 306-317. Kalantzis G, Emfietzoglou D, Hadjidoukas P 2012 A unified spatio-temporal parallelization framework for ac- celerated Monte Carlo radiobiological modeling of elec- tron tracks and subsequent radiation chemistry, Com- puter Physics Communications 183, 1683. Karplus M and Petsko G A 1990 Molecular dynamics simulations in biology, Nature 347, 631-639. Karplus M and McCammon J A 2002 Molecular dy- namics simulations of biomolecules, Nature Structural Biology 9, 646-652. McNamara A L, Guatelli S, Prokopovich D A, Rein- hard M I and Rosenfeld A. B. 2012 A comparison of X-ray and proton beam low energy secondary electron track structures using the low energy models of Geant4, International Journal of Radiation Biology, 88, 164-170. Michalik V 1993 Rad. Res. 134, 265-270; Verhaegena E, and Renier B 2004 Rad. Res. 162, 592-599. Moiseenko V V, Hamm R N, Waker A J, and Prestwich W V 1998 Modelling DNA damage induced by different energy photons and tritium beta-particles, Int. J. Radiat. Biol. 74, 533-550. Monti S, van Duin A C T, Kim S-Y, Barone V 2012, Ex- ploration of the Conformational and Reactive Dynamics of Glycine and Diglycine on TiO2: Computational In- vestigations in the Gas Phase and in Solution, J. Phys. Chem. C 116, 5141-5150. Mozumder A 1985 Early Production of Radicals from Charged Particle Tracks in Water, Radiat. Res. 104, 33-39. Munteanu M G, Vlahovicek K, Parthasaraty S, Simon I and Pongor S 1998 Rod models of DNA: sequence- dependent anisotropic elastic modeling of local bend- ing phenomena, Trends Biochem. Sci. 23, 341-346; http://hydra.icgeb.trieste.it/dna/model it.html Nikjoo H, and Charlton D E, 1995 Calculation of range and distributions of damage to DNA by high- and low- LET radiations in Radiation Damage to DNA: Struc- ture/Function Relationships at Early Times (A. F. Fu- ciarelli and J. D. Zimbrick, Eds. Battelle Press, Colum- bus). Nikjoo H, O'Neill P, Terrissol M, and Goodhead D T 1994 Modeling of radiation-induced DNA damage: the early physical and chemical event, Int. J. Radiat. Biol. 66, 453-457. Nikjoo H, O'Neill P, Goodhead D T and Terrissol M 1997 Computational modelling of low-energy electron- induced DNA damage by early physical and chemical events, Int. J. of Radiat. Biol. 71, 467-483. Paganetti H, Niemierko A, Ancukiewicz M, Gerweck L E, Goitein M, Loeffler J S, Suit H D 2003 Int J Radiat. Oncol. Biol. Phys., 53, 407. Plimpton S J 1995 J. Comp. Phys. 117, 1. Plimpton S J, Pollock R, Stevens M; Ewald P -M, 1997 in Proc of the Eighth SIAM Conference on Parallel Pro- cessing for Scientific Computing, Minneapolis, MN. Roots R and Okada S 1975 Estimation of life times and diffusion distances of radicals involved in x-ray-induced DNA strand breaks of killing of mammalian cells, Radiat Res. 64, 306-320. Schwartz S D and Schramm V L 2009 Enzymatic tran- sition states and dynamic motion in barrier crossing, Na- ture Chemical Biology 5, 551-558. Semenenko V A, Stewart R D, Ackerman E J, 2005 Monte Carlo Simulation of Base and Nucleotide Excision 10 Repair of Clustered DNA Damage Sites. I. Model Prop- erties and Predicted Trends, Radiat. Res. 164, 180-193; ibid. 2005 Comparisons of Model Predictions to Mea- sured Data.164, 194-201. Stewart R D, Yu V K, Georgakilas A G, Koumenis C, Park J H and Carlson D J 2011 Effects of Radiation Quality and Oxygen on Clustered DNA Lesions and Cell Death, Radiat. Res., 176, 587-602. Schneider G and Fechner U, 2005 Computer-based de novo design of drug-like molecules, Nature Reviews Drug Discovery 4, 649-663; Durrant J D and McCammon J A, 2011 Molecular dynamics simulations and drug discovery BMC Biology 9, 71. Sies H 1993 Strategies of antioxidant defense, Europ. J. Biochem. 215, 213. Terrisol M, and Beaudre A 1990 Simulation of space and time evolution of radiolytic species induced by elec- trons in water, Radiat. Prot. Dosimetry 31, 171. Van Der Spoel D, Lindahl E, Hess B, Groenhof G, Mark A E, Berendsen H J C 2005 J. Comput. Chem. 26, 1701-1718. van Duin A C T, Dasgupta S, Lorant F, Goddard W A 2001 J. Phys. Chem. A 105, 9396-9409. Wilson W E, and Paretzke H G 1994 A stochastic model of ion track structure, Radiat. Prot. Dosimetry 52, 249. Ward J F 1988 DNA damage produced by ionizing radi- ation in mammalian cells: identities, mechanisms of for- mation, and reparability, Prog. Nucleic Acid Res. Mol. Biol. 35 95-125. Zaider M, and Brenner D J 1984 On the stochastic treatment of fast chemical reactions, Radiat. Res., 100, 245-256.
1803.03201
2
1803
2018-10-22T14:12:31
Quantifying the impact of a periodic presence of antimicrobial on resistance evolution in a homogeneous microbial population of fixed size
[ "physics.bio-ph", "q-bio.PE" ]
The evolution of antimicrobial resistance generally occurs in an environment where antimicrobial concentration is variable, which has dramatic consequences on the microorganisms' fitness landscape, and thus on the evolution of resistance. We investigate the effect of these time-varying patterns of selection within a stochastic model. We consider a homogeneous microbial population of fixed size subjected to periodic alternations of phases of absence and presence of an antimicrobial that stops growth. Combining analytical approaches and stochastic simulations, we quantify how the time necessary for fit resistant bacteria to take over the microbial population depends on the alternation period. We demonstrate that fast alternations strongly accelerate the evolution of resistance, reaching a plateau for sufficiently small periods. Furthermore, this acceleration is stronger in larger populations. For asymmetric alternations, featuring a different duration of the phases with and without antimicrobial, we shed light on the existence of a minimum for the time taken by the population to fully evolve resistance. The corresponding dramatic acceleration of the evolution of antimicrobial resistance likely occurs in realistic situations, and may have an important impact both in clinical and experimental situations.
physics.bio-ph
physics
Quantifying the impact of a periodic presence of antimicrobial on resistance evolution in a homogeneous microbial population of fixed size Loıc Marrec1, Anne-Florence Bitbol1* 1 Sorbonne Universit´e, CNRS, Laboratoire Jean Perrin (UMR 8237), F-75005 Paris, France * [email protected] Abstract The evolution of antimicrobial resistance generally occurs in an environment where antimicro- bial concentration is variable, which has dramatic consequences on the microorganisms' fitness landscape, and thus on the evolution of resistance. We investigate the effect of these time-varying patterns of selection within a stochastic model. We consider a homogeneous microbial population of fixed size subjected to periodic alternations of phases of absence and presence of an antimicrobial that stops growth. Combining analyti- cal approaches and stochastic simulations, we quantify how the time necessary for fit resistant bacteria to take over the microbial population depends on the alternation period. We demon- strate that fast alternations strongly accelerate the evolution of resistance, reaching a plateau for sufficiently small periods. Furthermore, this acceleration is stronger in larger populations. For asymmetric alternations, featuring a different duration of the phases with and without an- timicrobial, we shed light on the existence of a minimum for the time taken by the population to fully evolve resistance. The corresponding dramatic acceleration of the evolution of antimi- crobial resistance likely occurs in realistic situations, and may have an important impact both in clinical and experimental situations. Introduction The discovery of antibiotics and antivirals has constituted one of the greatest medical advances of the twentieth century, allowing many major infectious diseases to be treated. However, with the increasing use of antimicrobials, pathogenic microorganisms tend to become resistant to these drugs. Antimicrobial resistance has become a major and urgent problem of public health worldwide [1, 2]. Mutations that confer antimicrobial resistance are often associated with a fitness cost, i.e. a slower reproduction [3 -- 5]. Indeed, the acquisition of resistance generally involves either a mod- ification of the molecular target of the antimicrobial, which often alters its biological function, or the production of specific proteins, which entails a metabolic cost [4]. However, resistant microorganisms frequently acquire subsequent mutations that compensate for the initial cost of resistance. These microorganisms are called "resistant-compensated" [6 -- 9]. The acquisi- tion of resistance is therefore often irreversible, even if the antimicrobial is removed from the environment [4, 6]. In the absence of antimicrobial, the adaptive landscape of the microorganism, which repre- sents its fitness (i.e. its reproduction rate) as a function of its genotype, involves a valley, since the first resistance mutation decreases fitness, while compensatory mutations increase it. How- ever, this fitness valley, which exists in the absence of antimicrobial, disappears above a certain concentration of antimicrobial, as the growth of the antimicrobial-sensitive microorganism is impaired. Thus, the adaptive landscape of the microorganism depends drastically on whether the antimicrobial is present or absent. Taking into account this type of interaction between genotype and environment constitutes a fundamental problem, even though most experiments have traditionally focused on comparing different mutants in a unique environment [10]. In par- ticular, recent theoretical analyses show that variable adaptive landscapes can have a dramatic evolutionary impact [11 -- 15]. 1 How do the timescales of evolution and variation in the adaptive landscape compare and interact? What is the impact of the time variability of the adaptive landscape on the evolution of antimicrobial resistance? In order to answer these questions, we construct a minimal model retaining the fundamental aspects of antimicrobial resistance evolution. Focusing on the case of a homogeneous microbial population of fixed size, we perform a complete stochastic study of de novo resistance acquisition in the presence of periodic alternations of phases of absence and presence of an antimicrobial that stops growth. These alternations can represent, for example, a treatment where the concentration within the patient falls under the Minimum Inhibitory Con- centration (MIC) between drug intakes [16]. Combining analytical and numerical approaches, we show that these alternations substantially accelerate the evolution of resistance with respect to the cases of continuous absence or continuous presence of antimicrobial, especially for larger populations. We fully quantify this effect and shed light on the different regimes at play. For asymmetric alternations, featuring a different duration of the phases with and without antimi- crobial, we demonstrate the existence of a minimum for the time taken by the population to fully evolve resistance, occurring when both phases have durations of the same order. This realistic situation dramatically accelerates the evolution of resistance. Finally, we discuss the implications of our findings, in particular regarding antimicrobial dosage. Model The action of an antimicrobial drug can be quantified by its MIC, the minimum concentration that stops the growth of a microbial population [4]. We focus on biostatic antimicrobials, which stop microbial growth (vs. biocidal antimicrobials, which kill microorganisms). We model the action of the antimicrobial in a binary way: below the MIC ("absence of antimicrobial"), growth is not affected, while above it ("presence of antimicrobial"), sensitive microorganisms cannot grow at all. The usual steepness of pharmacodynamic curves around the MIC [16] justifies our simple binary approximation, and we also present an analysis of the robustness of this hypothesis (Supplementary Material, Section 6). Within this binary approximation, there are two adaptive landscapes. Assuming that the drug fully stops the growth of sensitive microorganisms, but does not affect that of resistant ones, and considering compensatory mutations that fully restore fitness, these two adaptive landscapes can be described by a single parameter δ, representing the fitness cost of resistance (Fig. 1A). We focus on asexual microorganisms, and fitness simply denotes the division rate of these organisms. The fitness of sensitive microorganisms in the absence of antimicrobials is taken as reference. In this framework, we investigate the impact of a periodic presence of antimicrobial, assuming that the process starts without antimicrobial (Fig. 1B-C). 1 0 1 2 1 0 Antimicrobial presence Time 0 T 2T Figure 1: Model. (A) Adaptive landscapes in the presence and in the absence of antimicrobial. Genotypes are indicated by the number of mutations from the sensitive microorganism, and by initials: S: sensitive; R: resistant; C: resistant-compensated. (B) and (C) Periodic presence of antimicrobial, and impact on the fitness of S (sensitive) mi- croorganisms: (B) Symmetric alternations; (C) Asymmetric alternations. We denote by µ1 and µ2 the mutation rates (or mutation probabilities upon each division) 2 for the mutation from S to R and for the one from R to C, respectively. In several actual situations, the effective mutation rate towards compensation tends to be higher than the one towards the return to sensitivity, since multiple mutations can compensate for the initial cost of resistance [7, 8, 17]. Therefore, we do not take into account back-mutations. Still because of the abundance of possible compensatory mutations, generally µ1 ≪ µ2 [7, 18]. We present general analytical results as a function of µ1 and µ2, and analyze in more detail the limit µ1 ≪ µ2, especially in simulations. All notations introduced are summed up in Table S1. We focus on a homogeneous microbial population of fixed size N , which can thus be described in the framework of the Moran process [19, 20], where fitnesses are relative (see Supplementary Material, Section 2 and Fig. S1). Assuming a constant size simplifies the analytical treatment and is appropriate for instance to describe turbidostat experiments, where the dilution rate is adjusted so that turbidity (and hence population size) is constant [21]. If a population only features sensitive individuals (with zero fitness) in the presence of antimicrobial, we consider that no division occurs, and the population remains static. We always express time in number of generations, which corresponds (unless no cell can divide) to the number of Moran steps divided by the population size N . Throughout, we start from a microbial population where all individuals are S (sensitive), and C it takes for the C (resistant-compensated) type to fix in the population, we focus on the time tf i.e. to take over the population. Then, the population has fully evolved resistance de novo. Results A periodic presence of antimicrobial can drive resistance evolution In this section, we study how alternations of absence and presence of antimicrobial can drive the de novo evolution of resistance. We present analytical predictions for the time needed for the population to evolve resistance, and then we compare them to numerical simulation results. We first focus on the rare mutation regime N µ1 ≪ 1, where at most one mutant lineage exists in the population at each given time. The frequent mutation regime is briefly discussed, and more detail regarding the appropriate deterministic treatment in this regime is given in Supplementary Material, Section 3. Here, we consider the case of symmetric alternations with period T (Fig. 1B). Asymmetric alternations (Fig. 1C) will be discussed later. Time needed for resistant microorganisms to start growing Resistant (R) mutants can only appear during phases without antimicrobial. Indeed, mutations occur upon division, and sensitive (S) bacteria cannot divide in the presence of antimicrobial (Fig. 1). However, R mutants are less fit than S individuals without antimicrobial. Hence, the lineage of an R mutant will very likely disappear, unless it survives until the next addition of antimicrobial. More precisely, without antimicrobial, the fixation probability pSR of a single R mutant with fitness 1 − δ, in a population of size N where all other individuals are of type S and have fitness 1, is ∼ 1/N if the mutation from S to R is effectively neutral (N δ ≪ 1), and ∼ δe−N δ if δ ≪ 1 and N δ ≫ 1 [20]. Let us denote by τ d R the average time an R lineage would drift before going extinct without antimicrobial [20] (see Supplementary Material, Section 2). If antimicrobial is added while R mutants exist in the population, i.e. within ∼ τ d R after a mutation event, then the R population will grow fast and fix, since S individuals cannot divide with antimicrobial. Hence, each time antimicrobial is added, any R lineage that was destined for extinction without antimicrobial but that survived until the addition of drug is rescued. Through this phenomenon, periodic alternations of absence and presence of antimicrobial can substantially accelerate resistance evolution: we will quantify this effect. Note that here, we disregard the very few R lineages destined for fixation without antimicrobial, because we aim to study the acceleration of resistance evolution due to the alternations. The spontaneous evolution of resistance without antimicrobial is discussed and compared to our alternation-driven process in the Supplementary Material, Section 4. 3 It is crucial to calculate the average waiting time ta R until an R lineage is rescued by the addition of antimicrobial. Indeed, this constitutes the key step of alternation-driven resistance takeover. Three timescales impact ta R. The first one is the timescale of the environment, namely the half-period T /2. The two other ones are intrinsic timescales of the evolution of the population without antimicrobial: the average time between the appearance of two independent R mutants, 1/(N µ1), and the average lifetime τ d R of the lineage of an R mutant destined for extinction without antimicrobial. Note that τ d R ≈ log N for large R is generally quite short. N if δ = 0, and τ d R decreases as δ increases, as deleterious R mutants are out-competed by S microorganisms; for instance, τ d R ≈ 2.6 generations if δ = 0.1 in the limit where N ≫ 1 and N δ ≫ 1 [20] (see Supplementary Information, Section 2). Hence, in the rare mutation regime, τ d R ≪ 1/(N µ1). What matters is how the environment timescale T /2 compares to these two evolution timescales (see Fig. 2A-C). Our arguments based on comparing average timescales are approximate, but they yield explicit analytical predictions in each regime where timescales are separated, which we then test through numerical simulations. Indeed, τ d R, where τ d Figure 2: Alternation-driven evolution of antimicrobial resistance. (A-C) Sketches il- lustrating the three different regimes for the half-period T /2 of the alternations of antimicrobial absence (white) and presence (gray). The fraction of resistant (R) microorganisms in the pop- ulation is plotted versus time (blue curves). R mutants can only appear without antimicrobial. (A) T /2 ≪ τ d R is the average extinction time of the lineage of an R mutant without antimicrobial. The first R lineage that appears is expected to live until the next addition of R ≪ T /2 ≪ 1/(N µ1), where 1/(N µ1) is the aver- antimicrobial and is then rescued. age time between the appearance of two independent R mutants without antimicrobial. (C) T /2 ≫ 1/(N µ1). In (B) and (C), not all R lineages live until the next addition of antimicrobial, and in (C) multiple R lineages arise within a half-period. (D) Example of a simulation run. The fractions of S, R and C microorganisms are plotted versus time. Inset: end of the process, with full resistance evolution. As in (A-C), antimicrobial is present during the gray-shaded time intervals (shown only in the inset given their duration). Parameters: µ1 = 10−5, µ2 = 10−3, δ = 0.1, N = 102 and T = 50 (belonging to regime B). (B) τ d (A) If T /2 ≪ τ d exist upon the next addition of antimicrobial, and to be rescued, which yields ta Indeed, mutations from S to R can only occur without antimicrobial, i.e. half of the time. R (Fig. 2A): The lineage of the first R mutant that appears is likely to still R = 2/(N µ1). (B) If τ d R ≪ T /2 ≪ 1/(N µ1) (Fig. 2B): At most one mutation yielding an R individual is expected within each half-period. The lineage of this mutant is likely to survive until the next addition of antimicrobial only if the mutant appeared within the last ∼ τ d R preceding it, which has a probability p = 2τ d R = 2/(N µ1p) = T /(N µ1τ d R/T . Hence, ta R). 4 (C) If T /2 ≫ 1/(N µ1) (Fig. 2C): Since the half-period is much larger than the time 1/(N µ1) between the appearance of two independent mutants without antimicrobial, several appearances and extinctions of R lineages are expected within one half-period. Hence, the probability that a lineage of R exists upon a given addition of antimicrobial is q = N µ1τ d R, which corresponds to the fraction of time during which R mutants are present in the phases without antimicrobial. Specifically, q is the ratio of the average lifetime of the lineage of an R mutant destined for extinction without antimicrobial to the average time between the appearance of two independent R mutants without antimicrobial. Since additions of antimicrobial occur every T , we have ta R = T /q = T /(N µ1τ d R), which is the same as in case (B). In fact, the demonstration presented for case (C) also holds for case (B). In conclusion, we obtain ta R = T N µ1 min(cid:0)τ d R, T /2(cid:1) . Hence, if T /2 ≪ τ d is proportional to T . R, ta R is independent from the period T of alternations, while if T /2 ≫ τ d R, ta R Time needed for the population to fully evolve resistance We are interested in the average time tf C it takes for the population to fully evolve resistance, i.e. for the C (resistant-compensated) type to fix. An example of the process is shown in Fig. 2D. It takes on average ta R for R mutants to be rescued by the addition of antimicrobial. Then they rapidly grow, since S individuals cannot divide. If the phase with antimicrobial is long enough, R mutants take over during this phase, with a probability 1 and an average fixation timescale τ f R ≈ log N for N ≫ 1 [20] (see Supplementary Material, Section 2). If T /2 ≪ τ f R, fixation cannot occur within a single half-period, and the R lineage will drift longer, but its extinction remains very unlikely. Indeed, while R individuals are the only ones that can divide with antimicrobial, we assume that they experience only a minor disadvantage without antimicrobial (1 − δ vs. 1, generally with δ ≪ 1 [4], see Fig. 1A). Hence, if T /2 ≪ τ f R, and neglecting changes in frequencies in the absence of antimicrobial, R mutants will take ∼ 2τ f R to fix. Once the R type has fixed in the population, the appearance and eventual fixation of C mutants are independent from the presence of antimicrobial, since only S microorganisms are affected by it (see Fig. 1A). The first C mutant whose lineage will fix takes an average time ta C = 1/(N µ2 pRC) to appear once R has fixed, where pRC is the fixation probability of a single C mutant in a population of size N where all other individuals are of type R. In particular, if N δ ≪ 1 then pRC = 1/N , and if δ ≪ 1 and N δ ≫ 1 then pRC ≈ δ [20] (see Supplementary Material, Section 2). The final step is the fixation of this successful C mutant, which will take an average time τ f C, of order N in the effectively neutral regime N δ ≪ 1, and shorter for larger δ given the selective advantage of C over R [20] (see Supplementary Material, Section 2). Note that we have assumed for simplicity that the fixation of R occurs before the appearance of the first successful C mutant, which is true if ta R, i.e. 1/(N µ2 pRC) ≫ log N . This condition is satisfied if the second mutation is sufficiently rare. Otherwise, our calculation will slightly overestimate the actual result. C ≫ τ f Combining the previous results yields R + ta R + τ f tf C ≈ ta C = 1/(N µ2 pRC), and τ f C + τ f C , R is given by Eq. 1, while ta where ta mutation regime, the contribution of the two fixation times τ f addition µ1 ≪ µ2, which is realistic (cf. Methods), then tf tf C will be dominated by ta that the condition ta R > ta Hence, T > 2τ d is equivalent to T > τ d ta C to tf C . N . In the rare C will be negligible. If in R. If µ1 ≈ µ2, R if T > max(cid:0)2τ d R, using Eq. 1 shows C is then equivalent to pRC > 1/2, which cannot be satisfied for δ ≪ 1. R > ta R/pRC(cid:1), the contribution of R/pRC. Beyond the regime T > max(cid:0)2τ d C R and µ1 ≈ µ2, the condition ta R and τ f R ≈ log N and τ f R/pRC(cid:1). Indeed, if T < 2τ d C will be dominated by ta R is necessary to have ta C will be important. C. But if T > 2τ d R > ta R, τ d R, τ d (1) (2) 5 Comparison of analytical predictions and simulation results R, tf R. If T ≪ 2τ d C does not depend on T , while if T ≫ 2τ d Fig. 3A shows simulation results for the average total fixation time tf C of C individuals in the population. This time is plotted as a function of the period T of alternations for different population sizes N . As predicted above (see Eq. 1), we observe two regimes delimited by T = 2τ d R, it depends linearly on T . In Fig. 3A, we also plot our analytical prediction from Eqs. 1 and 2 in these two regimes (solid lines). The agreement with our simulated data is excellent for small and intermediate values of T , without any adjustable parameter. Interestingly, the transition between these two regimes occurs for periods of about 5 generations, which would correspond to a few hours for typical bacteria, thus highlighting the practical importance of these two regimes. In Fig. 3A, the smallest values reported for tf C are of order 100 generations, corresponding to a few days, and are thus relevant to an actual treatment, while some other values are larger than the timescales involved in a treatment. Here, we quantitatively analyze the phenomena for a wide range of parameters. A more detailed comparison to actual situations, employing realistic values of population sizes and mutation rates, is presented in the Discussion. Figure 3: Impact of symmetric alternations. Fixation time tf C of C (resistant-compensated) individuals in a population of N individuals subjected to symmetric alternations of absence and presence of antimicrobial with period T . Data points correspond to the average of simulation results, and error bars (often smaller than markers) represent 95% confidence intervals. 2 to 104 replicate simulations were performed in each case (the smallest numbers of replicates were used for the largest populations, whose evolution is quasi-deterministic). In both panels, solid lines correspond to our analytical predictions in each regime. Parameter values: µ1 = 10−5, µ2 = 10−3, and δ = 0.1. (B) tf C limit of the neutral regime, N = 1/δ. Right as function of N . Left vertical dashed line: vertical dashed line: limit of the deterministic regime, N = 1/µ1. Horizontal purple line: analytical prediction for valley crossing by neutral tunneling in the presence of alternations (see Supplementary Material, Section 4). Black lines: analytical predictions for fitness valley crossing times in the absence of alternations (see Supplementary Material, Section 4). C as function of T . Vertical dashed line: T = 2τ d R. (A) tf Importantly, Fig. 3A shows that tf C reaches a plateau for small N and large T , which is not predicted by our analysis of the alternation-driven evolution of resistance. This plateau corresponds to the spontaneous fitness valley crossing process [22], through which resistance mutations appear and fix in the absence of drug. Note that such a plateau would also be reached for larger N , but for periods T longer than those considered in Fig. 3A (see Fig. 3B, black lines). What ultimately matters is the shortest process among the alternation-driven one and the spontaneous valley-crossing one. In Fig. 3A, horizontal solid lines at large T represent our analytical predictions for the valley-crossing time (see Supplementary Material, Section 4). C as function of N for different T . Again, solid lines rep- Fig. 3B shows simulation results for tf 6 resent our analytical predictions from Eqs. 1 and 2, yielding excellent agreement for intermediate values of N , and for small ones at small T . In other regimes, resistance evolution is achieved by spontaneous valley crossing. In the limit T → ∞ of continuous absence of antimicrobial (black data points in Fig. 3B), only valley crossing can occur, and the black solid lines correspond to our analytical predictions for this process (see Supplementary Material, Section 4). Until now, we focused on the rare mutation regime. In the large-population, frequent- mutation regime N ≫ 1/µ1 ≫ 1, the dynamics of the population can be well-approximated by a deterministic model with replicator-mutator differential equations [23, 24] (see Supplemen- tary Material, Section 3). Then, several lineages of mutants can coexist. If T /2 ≫ 1/(N µ1), it is almost certain that some R mutants exist in the population upon the first addition of antimicrobial, which entails ta R = T /2. The horizontal purple solid line plotted at large T in Fig. 3A, and the horizontal solid lines at large N in Fig. 3B, both correspond to this determinis- tic prediction. In the Supplementary Material, Section 3, we study the deterministic limit of our stochastic model, and demonstrate that it matches the results obtained in Fig. 3A for N = 105 and N = 106 over the whole range of T (see Fig. S2). The comparison to the spontaneous fitness valley crossing process (Fig. 3B, black curve and Supplementary Material, Section 4) demonstrates that periodic alternations of absence and pres- ence of antimicrobial can dramatically accelerate resistance evolution compared to continuous absence of antimicrobial. Recall that within our model, sensitive microorganisms cannot divide with antimicrobial, so resistance cannot evolve at all in continuous presence of antimicrobial. Another possible comparison would be to a continuous presence of a low dose of antimicrobial (below the MIC), but this goes beyond our binary model of antimicrobial action (see Supplemen- tary Information, Section 6 for a discussion of the domain of validity of this model). Alternations are really essential: R mutants appear without antimicrobial, and each addition of antimicrobial rescues the existing R lineages that would be destined to extinction without antimicrobial. Asymmetric alternations The average time ta We now turn to the more general case of asymmetric alternations of phases of absence and presence of antimicrobial, with respective durations T1 and T2, and T = T1 + T2 (see Fig. 1C). R when R mutants first exist in the presence of antimicrobial, and start growing, can be obtained by a straightforward generalization of the symmetric alternation case Eq. 1. What matters is how the duration T1 of the phase without antimicrobial, where S individuals can divide and mutate, compares to the average time τ d R an R lineage would drift before extinction without antimicrobial. R, the first R mutant takes an average time T /(N µ1T1) to appear, and is likely to be rescued by the next addition of antimicrobial. If T1 ≫ τ d R, the fraction of time during which R mutants are present in the phases without antimicrobial is N µ1τ d R). Hence, we obtain R, and antimicrobial is added every T , so ta If T1 ≪ τ d R = T /(N µ1τ d ta R = T N µ1 min(τ d R, T1) . (3) Once the R mutants have taken over the population, the appearance and fixation of C mutants is not affected by the alternations. Hence, Eq. 2 holds for asymmetric alternations, with ta R given by Eq. 3. In the rare mutation regime, if µ1 ≪ µ2, then tf C will be dominated by ta R, and if µ1 ≈ µ2, then tf fixation probability of a single C mutant in a population of R individuals. R, T1(cid:1) /pRC, where pRC is the R if T > min(cid:0)τ d C will be dominated by ta Fig. 4A shows simulation results for tf C as a function of the duration T1 of the phases without antimicrobial, for different values of the duration T2 of the phases with antimicrobial. As predicted above, we observe a transition at T1 = τ d R, and different behaviors depending whether T2 ≪ τ f R. Our analytical predictions from Eqs. 2 and 3 are plotted in Fig. 4A in the various regimes (solid lines), and are in excellent agreement with the simulation data. The plateau of tf C at large T1 corresponds to spontaneous valley crossing, and the analytical prediction (see Supplementary Material, Section 4) is plotted in black in Fig. 4A. R or T2 ≫ τ f 7 Figure 4: Asymmetric alternations. Fixation time tf C of C individuals in a population sub- jected to asymmetric alternations of absence and presence of antimicrobial (respective durations: T1 and T2). Data points correspond to the average of simulation results (over 10 to 103 repli- cates), and error bars (sometimes smaller than markers) represent 95% confidence intervals. In both panels, solid lines correspond to our analytical predictions in each regime. In particular, black lines are analytical predictions for fitness valley crossing times in the absence of alterna- tions (see Supplementary Material, Section 4). Parameter values: µ1 = 10−5, µ2 = 10−3, δ = 0.1 and N = 103. (A) tf C as function of T2 for different T1. Dashed line: T2 = τ f R. C as function of T1 for different T2. Dashed line: T1 = τ d R. (B) tf For T2 ≫ τ f R, Fig. 4A shows that tf C features a striking minimum, which gets higher but wider for longer T2. This can be fully understood from our analytical predictions. Indeed, when T1 is varied starting from small values at fixed T2 ≫ τ f R, different regimes can be distinguished: R(cid:0). τ f R ≪ T2(cid:1), Eq. 3 yields ta R = T /(N µ1T1) ≈ T2/(N µ1T1) ∝ 1/T1. R ≪ T1 ≪ T2, Eq. 3 gives ta R = T /(N µ1τ d R) ≈ T2/(N µ1τ d R), which is independent • When T1 ≪ τ d • When τ d from T1. • As T1 reaches and exceeds T2, the law ta R) ∝ T1 when τ d R ≪ T2 ≪ T1. T1/(N µ1τ d R = T /(N µ1τ d R) still holds. It yields ta R ≈ Hence, the minimum of ta but wider for larger T2. R is T2/(N µ1τ d In the opposite regime where T2 ≪ τ d as a function of T1: R) ∝ T2 and is attained for τ d R . τ f R, Fig. 4A shows that tf R ≪ T1 ≪ T2: it gets higher C also features a minimum • When T1 ≪ T2 ≪ τ d • When T2 ≪ T1 ≪ τ d R, Eq. 3 yields ta R, the same law gives ta R = T /(N µ1T1) ≈ T2/(N µ1T1) ∝ 1/T1. R = T /(N µ1T1) ≈ 1/(N µ1), which is indepen- dent from T1. • When T2 ≪ τ d R ≪ T1, R lineages eventually tend to go extinct, even once they have started growing thanks to an addition of antimicrobial (see Supplementary Material, Section 5 and Fig. S3B). Then, alternations do not accelerate resistance evolution, and spontaneous valley crossing dominates (black horizontal line in Fig 4A). Hence, the minimum of ta R is 1/(N µ1) and is attained for T2 ≪ T1 ≪ τ d R: then, the first R mutant that appears is likely to be rescued by the next addition of antimicrobial, thus driving R, ta the complete evolution of resistance in the population. For T2 ≤ T1 ≪ τ d R is between once and twice this minimum value. 8 A similar analysis can be conducted if T2 is varied at fixed T1 (Fig. 4B); it is presented in the Supplementary Material, Section 5. In a nutshell, for asymmetric alternations, a striking minimum for the time of full evolution of resistance by a population occurs when both phases have durations of the same order. Interestingly, the minimum generally occurs when the phases of antimicrobial presence are shorter than those of absence, i.e. T2 ≤ T1 (except if T2 ≫ τ d R). In addition to this minimum, Fig. 4 also shows a regime of parameters, when T1 ≪ T2 and T1 ≪ τ d R, where the evolution of resistance actually takes longer than fitness valley crossing in the absence of antimicrobial (black lines in Fig. 4). Comparing the timescales involved (see Supplementary Material, Section 4) shows that in this regime, if T2 ≫ T1δ/µ2, the alternation- driven process is faster than the valley-crossing process in the presence of alternations, and thus dominates, but it is slower than the valley-crossing process in the absence of antimicrobial. Hence, in this case, the drug actually slows down the evolution of resistance. Qualitatively, this is because the antimicrobial prevents mutants from arising when it is present. Discussion Main conclusions Because of the generic initial fitness cost of resistance mutations, alternations of phases of absence and presence of antimicrobial induce a dramatic time variability of the adaptive landscape associated to resistance evolution, which alternates back and forth from a fitness valley to an ascending landscape. Using a general and minimal theoretical model which retains the key biological ingredients, we have shed light on the quantitative implications of these time-varying patterns of selection on the time it takes for resistance to fully evolve de novo in a homogeneous microbial population of fixed size. Combining analytical approaches and simulations, we showed that resistance evolution can be driven by periodic alternations of phases of absence and presence of an antimicrobial that stops growth. Indeed, the addition of antimicrobial is able to rescue resistant lineages that were destined to go extinct without antimicrobial. We found that fast alternations strongly accelerate the evolution of resistance. In the limit of short alternation periods, the very first resistant mutant that appears is likely to ultimately lead to full resistance of the population, as it will generally be rescued by the next addition of antimicrobial before going extinct, which would be its most likely fate without antimicrobial. For larger periods T , the time needed for resistance to evolve increases linearly with T , until it reaches the spontaneous valley-crossing time with alternations, which constitutes an upper bound. Our complete stochastic model allowed us to investigate the impact of population size N , beyond the limit N ≫ 1/µ1 addressed by deterministic models. We showed that the acceleration of resistance evolution is stronger for larger populations, eventually reaching a plateau in the deterministic limit. Over a large range of intermediate parameters, the time needed for the population to fully evolve resistance scales as T /N . These results are summed up in Fig. 5A. For asymmetric alternations, featuring different durations T1 and T2 of the phases of absence and presence of antimicrobial, we have shed light on the existence of a minimum for the time taken by the population to fully evolve resistance. This striking minimum occurs when both phases have durations of the same order, generally with T1 ≤ T2. Moreover, the minimum value reached for the time of resistance evolution decreases for shorter alternation periods. These results are summed up in Fig. 5B. Context and perspectives Our approach is complementary to previous studies providing a detailed modeling of specific treatments [16, 25 -- 30]. Indeed, the majority of them [16, 25 -- 28, 31, 32] neglect stochastic effects, while they can have a crucial evolutionary impact [20, 33]. The deterministic approach is appropriate if the number N of competing microbes satisfies N µ1 ≫ 1, where µ1 is the mutation rate [33, 34]. Such large sizes can be reached in some established infections [17], but microbial populations go through very small bottleneck sizes (sometimes N ∼ 1−10 [35]) when an infection 9 Figure 5: Heatmaps. Fixation time tf C of C individuals in a population of size N subjected to periodic alternations of absence and presence of antimicrobial. Simulation data plotted in Figs. 3A and 4A are linearly interpolated. Parameter values: µ1 = 10−5, µ2 = 10−3, δ = 0.1. (A) Symmetric alternations: tf C as function of the period T and the population size N . Top horizontal line: deterministic regime limit N = 1/µ1. Bottom horizontal line: neutral regime limit N = 1/δ. Quasi-vertical curve: T = 2τ d R. Diagonal line: T = N . Note that no data is shown for T /2 < 1/N because of the discreteness of our model, which can only deal with timescales larger or equal to the duration of one Moran step, i.e. 1/N generation. (B) Asymmetric alternations: tf C as function of the durations T1 and T2 of the phases of absence and presence of antimicrobial. Vertical line: T1 = τ d R. Diagonal line: T1 = T2. Here N = 103, so the first resistant mutant appears after an average time T /(N µ1T1) = 102 T /T1. R. Horizontal line: T2 = τ f is transmitted. Moreover, established microbial populations are structured, even within a single patient [36], and competition is local, which decreases the effective value of N . Some previous studies did take stochasticity into account, but several did not include compensation of the cost of resistance [37, 38], while others made specific epidemiological assumptions [29]. Given the usual steepness of pharmacodynamic curves [16], we have modeled the action of a biostatic antimicrobial in a binary way, with no growth inhibition under the MIC and full growth inhibition of S microorganisms above it (see Model). An analysis of the robustness of this approximation is presented in the Supplementary Material, Section 6, showing that it is appropriate if the rise time, i.e. the time needed for the fitness of sensitive microorganisms to switch from a low value to a high value and vice-versa when antimicrobial is removed or added, is short enough (see Fig. S4). Qualitatively, if this rise time is shorter than the other environmental and evolutionary timescales at play, then the fitness versus time function is effectively binary. Our model assumes that the size of the microbial population remains constant. While this is realistic in some controlled experimental setups, e.g. turbidostats [21], microbial populations involved in infections tend to grow, starting from a small transmission bottleneck, and the aim of the antimicrobial treatment is to make them decrease in size and eventually go extinct. In the case of biostatic antimicrobials, which prevent bacteria from growing, populations can go extinct due to spontaneous and immune system-induced death. Our model with constant population size should however be qualitatively relevant at the beginning and middle stages of a treatment (i.e. sufficiently after transmission and before extinction). Constant population sizes facilitate analytical calculations, and allowed us to fully quantify the impact of a periodic presence of antimicrobial on resistance evolution, but it will be very interesting to extend our work to variable population sizes [13, 39, 40]. This would allow us to model biocidal antimicrobials, and to include effects such as antibiotic tolerance, which tend to precede resistance under intermittent antibiotic exposure [41]. Another exciting extension would be to incorporate spatial structure [42 -- 44] and environment heterogeneity, in particular drug concentration gradients. Indeed, static gradients 10 can strongly accelerate resistance evolution [45 -- 48], and one may ask how this effect combines with the temporal alternation-driven one investigated here. Besides, it would be interesting to compare the impact of periodic alternations to that of random switches of the environment [11 -- 15]. Implications for clinical and experimental situations The situation where the phases of absence and presence of antimicrobial have similar durations (T1 ≈ T2) yields a dramatic acceleration of resistance evolution, and is unfortunately clinically realistic. Indeed, a goal in treatment design is that the serum concentration of antimicrobial exceeds the MIC for at least 40 to 50% of the time [49], which implies that actual treatments may involve the alternations that most favor resistance evolution according to our results [16, 49]. Besides, bacteria divide on a timescale of about an hour (yielding a τ d R of order of a few hours), and antimicrobial is often taken every 8 to 12 hours in treatments by the oral route, so the alternation period does not last for many generations: this is close to our worst-case scenario of short symmetric periods. In this worst case scenario, full de novo resistance evolution can result from the appearance of the very first R mutant, which takes T /(N µ1T1). Indeed, its lineage is likely to be rescued by the next addition of antimicrobial. Under the conservative assumption that only one resistance mutation is accessible, taking µ1 ∼ 10−10, which is the typical mutation probability per nu- cleotide and per generation in Escherichia coli bacteria [50], and taking δ ∼ 0.1 [6], we find that this duration is less than a day (∼ 10− 20 generations) for N ∼ 109, and a few days for N ∼ 108, numbers that can be reached in infections [5, 17]. For such large populations, the fixation of the C (compensated) mutant will take more time, but once R is fixed (which takes ∼ 1 day after the appearance of the first R mutant), C is very likely to fix even if the treatment is stopped. This is due to the large number of compensatory mutations, which yields a much higher effective mutation rate toward compensation than toward reversion to sensitivity [7, 8, 17]. In addition, many mutations to resistance are often accessible, yielding higher effective µ1, e.g. µ1 ∼ 10−8 for rifampicin resistance in some wild isolates of E. coli [9], meaning that smaller populations can also quickly become resistant in the presence of alternations. Recall that we are only con- sidering de novo resistance evolution, without pre-existent resistant mutants, or other possible sources of resistance, such as horizontal gene transfer, which would further accelerate resistance acquisition. In summary, an antimicrobial concentration that drops below the MIC between each intake can dramatically favor de novo resistance evolution. More specifically, we showed that the worst case occurs when T1 ≤ T2, which would be the case if the antimicrobial concentration drops below the MIC relatively briefly before each new intake. Our results thus emphasize how important it is to control for such apparently innocuous cases, and constitute a striking argument in favor of the development of extended-release antimicrobial formulations [51]. While the parameter range that strongly accelerates resistance evolution should preferably be avoided in clinical situations, it could be tested and harnessed in evolution experiments. Again, these parameters are experimentally accessible. Controlled variations of antimicrobial concentration are already used experimentally, in particular in morbidostat experiments [52], where the population size is kept almost constant, which matches our model. In Ref. [52], a dramatic and reproducible evolution of resistance was observed in ∼ 20 days when periodically adjusting the drug concentration to constantly challenge E. coli bacteria. Given our results, it would be interesting to test whether resistance evolution could be made even faster by adding drug in a turbidostat with a fixed periodicity satisfying T1 ≤ T2 ≪ τ d R. Acknowledgments We thank Claude Loverdo, David J. Schwab and Raphael Voituriez for stimulating discussions. AFB also acknowledges the KITP Program on Evolution of Drug Resistance (KITP, Santa 11 Barbara, CA, 2014), which was supported in part by the National Science Foundation under Grant NSF PHY 17-48958. LM acknowledges funding by a graduate fellowship from EDPIF. References [1] World Health Organization. Antimicrobial resistance: global report on surveillance; 2014. [2] UK Review on Antimicrobial Resistance, chaired by Jim O'Neill, 2016;. [3] Borman AM, Paulous S, Clavel F. Resistance of human immunodeficiency virus type 1 to protease inhibitors: selection of resistance mutations in the presence and absence of the drug. J Gen Virol. 1996 Mar;77 ( Pt 3):419 -- 426. [4] Andersson DI, Hughes D. Antibiotic resistance and its cost: is it possible to reverse resis- tance? Nat Rev Microbiol. 2010;8:260 -- 271. [5] zur Wiesch PA, Kouyos R, Engelstadter J, Regoes RR, Bonhoeffer S. Population biolog- ical principles of drug-resistance evolution in infectious diseases. Lancet Infect Dis. 2011 Mar;11(3):236 -- 247. [6] Schrag SJ, Perrot V, Levin BR. Adaptation to the fitness cost of antibiotic resistance in E. coli . Proc R Soc Lond B. 1997;264:1287 -- 1291. [7] Levin BR, Perrot V, Walker N. Compensatory mutations, antibiotic resistance and the population genetics of adaptive evolution in bacteria. Genetics. 2000 Mar;154(3):985 -- 997. [8] Paulander W, Maisnier-Patin S, Andersson DI. Multiple mechanisms to ameliorate the fitness burden of mupirocin resistance in Salmonella typhimurium. Mol Microbiol. 2007 May;64(4):1038 -- 1048. [9] Moura de Sousa J, Sousa A, Bourgard C, Gordo I. Potential for adaptation overrides cost of resistance. Future Microbiol. 2015;10(9):1415 -- 1431. [10] Taute KM, Gude S, Nghe P, Tans SJ. Evolutionary constraints in variable environments, from proteins to networks. Trends Genet. 2014 May;30(5):192 -- 198. [11] Mustonen V, Lassig M. Molecular evolution under fitness fluctuations. Phys Rev Lett. 2008 Mar;100(10):108101. [12] Rivoire O, Leibler S. The Value of Information for Populations in Varying Environments. J Stat Phys. 2011;142:1124 -- 1166. [13] Melbinger A, Vergassola M. The Impact of Environmental Fluctuations on Evolutionary Fitness Functions. Sci Rep. 2015 Oct;5:15211. [14] Desponds J, Mora T, Walczak AM. Fluctuating fitness shapes the clone-size distribution of immune repertoires. Proc Natl Acad Sci USA. 2016 Jan;113(2):274 -- 279. [15] Wienand K, Frey E, Mobilia M. Evolution of a fluctuating population in a randomly switching environment. Phys Rev Lett. 2017 Oct;119(15):158301. [16] Regoes RR, Wiuff C, Zappala RM, Garner KN, Baquero F, Levin BR. Pharmacodynamic functions: a multiparameter approach to the design of antibiotic treatment regimens. An- timicrob Agents Chemother. 2004 Oct;48(10):3670 -- 3676. [17] Hughes D, Andersson DI. Evolutionary consequences of drug resistance: shared principles across diverse targets and organisms. Nat Rev Genet. 2015 Aug;16(8):459 -- 471. 12 [18] Poon A, Davis BH, Chao L. The coupon collector and the suppressor mutation: esti- mating the number of compensatory mutations by maximum likelihood. Genetics. 2005 Jul;170(3):1323 -- 1332. [19] Moran PAP. Random processes in genetics. Mathematical Proceedings of the Cambridge Philosophical Society. 1958;54(1):60 -- 71. [20] Ewens WJ. Mathematical Population Genetics. Springer-Verlag; 1979. [21] Myers J, Clark LB. Culture conditions and the development of the photosynthetic mech- II. An apparatus for the continuous culture of Chlorella. J Gen Physiol. 1944 anism: Nov;28(2):103 -- 112. [22] Weissman DB, Desai MM, Fisher DS, Feldman MW. The rate at which asexual populations cross fitness valleys. Theor Pop Biol. 2009;75:286 -- 300. [23] Traulsen A, Claussen JC, Hauert C. Coevolutionary dynamics: from finite to infinite populations. Phys Rev Lett. 2005 Dec;95(23):238701. [24] Traulsen A, Hauert C. Stochastic evolutionary game dynamics. In: Schuster HG, editor. Reviews of Nonlinear Dynamics and Complexity. vol. II. Wiley-VCH; 2009. . [25] Lipsitch M, Levin BR. The population dynamics of antimicrobial chemotherapy. Antimicrob Agents Chemother. 1997 Feb;41(2):363 -- 373. [26] Wahl LM, Nowak MA. Adherence and drug resistance: predictions for therapy outcome. Proc Biol Sci. 2000 Apr;267(1445):835 -- 843. [27] Wu Y, Saddler CA, Valckenborgh F, Tanaka MM. Dynamics of evolutionary rescue in changing environments and the emergence of antibiotic resistance. J Theor Biol. 2014 Jan;340:222 -- 231. [28] Meredith HR, Lopatkin AJ, Anderson DJ, You L. Bacterial temporal dynamics enable optimal design of antibiotic treatment. PLoS Comput Biol. 2015 Apr;11(4):e1004201. [29] Abel Zur Wiesch P, Kouyos R, Abel S, Viechtbauer W, Bonhoeffer S. Cycling em- pirical antibiotic therapy in hospitals: meta-analysis and models. PLoS Pathog. 2014 Jun;10(6):e1004225. [30] Ke R, Loverdo C, Qi H, Sun R, Lloyd-Smith JO. Rational Design and Adaptive Manage- ment of Combination Therapies for Hepatitis C Virus Infection. PLoS Comput Biol. 2015 Jun;11(6):e1004040. [31] Schulz zur Wiesch P, Engelstadter J, Bonhoeffer S. Compensation of fitness costs and reversibility of antibiotic resistance mutations. Antimicrob Agents Chemother. 2010 May;54(5):2085 -- 2095. [32] Bauer M, Graf IR, Ngampruetikorn V, Stephens GJ, Frey E. Exploiting ecology in drug pulse sequences in favour of population reduction. PLoS Comput Biol. 2017 Sep;13(9):e1005747. [33] Fisher DS. Evolutionary Dynamics. In: Bouchaud JP, M´ezard M, Dalibard J, editors. Les Houches, Session LXXXV, Complex Systems. Elsevier; 2007. . [34] Rouzine IM, Rodrigo A, Coffin JM. Transition between stochastic evolution and deter- ministic evolution in the presence of selection: general theory and application to virology. Microbiol Mol Biol Rev. 2001 Mar;65(1):151 -- 185. [35] Gutierrez S, Michalakis Y, Blanc S. Virus population bottlenecks during within-host pro- gression and host-to-host transmission. Curr Opin Virol. 2012 Oct;2(5):546 -- 555. 13 [36] van Marle G, Gill MJ, Kolodka D, McManus L, Grant T, Church DL. Compartmentalization of the gut viral reservoir in HIV-1 infected patients. Retrovirology. 2007;4:87. [37] Nissen-Meyer S. Analysis of effects of antibiotics on bacteria by means of stochastic models. Biometrics. 1966;22(4):761 -- 780. [38] Hansen E, Woods RJ, Read AF. How to Use a Chemotherapeutic Agent When Resistance to It Threatens the Patient. PLoS Biol. 2017 Feb;15(2):e2001110. [39] Melbinger A, Cremer J, Frey E. Evolutionary game theory in growing populations. Phys Rev Lett. 2010 Oct;105(17):178101. [40] Huang W, Hauert C, Traulsen A. Stochastic game dynamics under demographic fluctua- tions. Proc Natl Acad Sci USA. 2015 Jul;112(29):9064 -- 9069. [41] Levin-Reisman I, Ronin I, Gefen O, Braniss I, Shoresh N, Balaban NQ. Antibiotic tolerance facilitates the evolution of resistance. Science. 2017 02;355(6327):826 -- 830. [42] Bitbol AF, Schwab DJ. Quantifying the role of population subdivision in evolution on rugged fitness landscapes. PLoS Comput Biol. 2014 Aug;10(8):e1003778. [43] Nahum JR, Godfrey-Smith P, Harding BN, Marcus JH, Carlson-Stevermer J, Kerr B. A tortoise-hare pattern seen in adapting structured and unstructured populations suggests a rugged fitness landscape in bacteria. Proc Natl Acad Sci USA. 2015 Jun;112(24):7530 -- 7535. [44] Cooper JD, Neuhauser C, Dean AM, Kerr B. Tipping the mutation-selection balance: Limited migration increases the frequency of deleterious mutants. J Theor Biol. 2015 Sep;380:123 -- 133. [45] Zhang Q, Lambert G, Liao D, Kim H, Robin K, Tung C, et al. Acceleration of emergence of bacterial antibiotic resistance in connected microenvironments. Science. 2011;333(6050):1764 -- 1767. [46] Greulich P, Waclaw B, Allen RJ. Mutational pathway determines whether drug gradients accelerate evolution of drug-resistant cells. Phys Rev Lett. 2012;109:088101. [47] Hermsen R, Deris JB, Hwa T. On the rapidity of antibiotic resistance evolution facilitated by a concentration gradient. Proc Natl Acad Sci USA. 2012;109:10775 -- 10780. [48] Baym M, Lieberman TD, Kelsic ED, Chait R, Gross R, Yelin I, et al. Spatiotemporal microbial evolution on antibiotic landscapes. Science. 2016 09;353(6304):1147 -- 1151. [49] Jacobs MR. Optimisation of antimicrobial therapy using pharmacokinetic and pharmaco- dynamic parameters. Clin Microbiol Infect. 2001 Nov;7(11):589 -- 596. [50] Wielgoss S, Barrick JE, Tenaillon O, Cruveiller S, Chane-Woon-Ming B, M´edigue C, et al. Mutation rate dynamics in a bacterial population reflect tension between adaptation and genetic load. G3. 2011;1:183 -- 186. [51] Gao P, Nie X, Zou M, Shi Y, Cheng G. Recent advances in materials for extended-release antibiotic delivery system. J Antibiot. 2011 Sep;64(9):625 -- 634. [52] Toprak E, Veres A, Michel JB, Chait R, Hartl DL, Kishony R. Evolutionary paths to antibi- otic resistance under dynamically sustained drug selection. Nat Genet. 2011 Dec;44(1):101 -- 105. [53] Taylor C, Iwasa Y, Nowak A. A symmetry of fixation times in evoultionary dynamics. Journal of Theoretical Biology. 2006;243:245 -- 251. [54] Sekimoto K. Stochastic Energetics. Springer-Verlag; 2010. 14 [55] Gardiner CW. Handbook of Stochastic Methods for Physics, Chemistry and the Natural Sciences. Springer; 1985. [56] Nowak MA, Komarova NL, Sengupta A, Jallepalli PV, Shih IM, Vogelstein B, et al. The role of chromosomal instability in tumor initiation. Proc Natl Acad Sci USA. 2002 Dec;99(25):16226 -- 16231. [57] Weinreich DM, Chao L. Rapid evolutionary escape in large populations from local peaks on the Wrightian fitness landscape. Evolution. 2005;59:1175 -- 1182. [58] Weissman DB, Feldman MW, Fisher DS. The rate of fitness-valley crossing in sexual populations. Genetics. 2010;186:1389 -- 1410. 15 Supplementary Material 1 Table of notations Notation Definition S R C T T1 T2 N δ µ1 µ2 tf C Sensitive microorganisms Resistant microorganisms Resistant-compensated microorganisms Period of the alternations of absence and presence of antimicrobial Duration of the phase without antimicrobial (for asymmetric alternations) Duration of the phase with antimicrobial (for asymmetric alternations) Population size Fitness cost of antimicrobial resistance Mutation rate from S to R Mutation rate from R to C Total time of full resistance evolution (time until the C type fixes, starting from a population of S individuals) Average time when R individuals first exist in the presence of antimicrobial, start- ing from a population of S individuals Average time when the first C mutant whose lineage will fix appears, starting from a population of R individuals Average lifetime of the lineage of a single R mutant, until it disappears, in a population of S individuals, in the absence of antimicrobial Average fixation time of the lineage of a single R mutant in a population of S individuals, in the presence of antimicrobial Average fixation time of the lineage of a single C mutant in a population of R individuals Fixation probability of a single R mutant in a population of S individuals in the absence of antimicrobial Fixation probability of a single C mutant in a population of R individuals ta R ta C τ d R τ f R τ f C pSR pRC Table S1: Notations. This table lists the different notations introduced in the main text and their meaning. 2 Fixation probabilities and fixation times in the Moran process Here, we discuss in detail the fixation probabilities and mean fixation times in the Moran process, which are used throughout the main text. These quantities are already known [20, 24], but we present a derivation for the sake of pedagogy and completeness. Our derivation is based on the general formalism of first passage times, and gives the same results as those obtained in the literature, often using other methods [20, 24]. Next, we use the general expressions obtained to express the various fixation probabilities and fixation times used in the main text. 2.1 The Moran process The Moran model [19, 20] is a simple stochastic process used to describe the evolution of the composition of asexual populations of finite and constant size. It allows one to incorporate variety-increasing processes such as mutation and variety-reducing processes such as natural selection. 16 In the Moran model, at each time step, an individual is chosen at random to reproduce and another one is chosen to die (see Figure S1). Hence, the total number of individuals in the population stays constant. Note that we will consider that the same individual can be selected to reproduce and die at the same step. Natural selection can be introduced by choosing the individual that reproduces with a probability proportional to its fitness. To implement mutations upon division, one can allow the offspring to switch type with a certain probability at each step. When a mutant arises within the Moran model at constant fitness, its lineage can either disappear or fix in the population, i.e. take over the whole population. The outcome is not fully determined by fitness differences as in a deterministic case, but also by stochastic fluctuations, also known as genetic drift. Here, we focus on the evolution of population composition under genetic drift and selection alone. In the rare mutation regime, these processes are much faster than the time between the occurrence of two mutations, so mutation can be neglected during the process of fixation of one type. The Moran model allows us to compute explicit expressions for quantities such as fixation probabilities and fixation times [20, 53] (see below). Figure S1: Sketch of the Moran process. One step of the Moran process is represented in a population with 8 individuals of 2 different types (different colors). Let us consider a population of N individuals of two types A and B, which have fitnesses fA and fB, respectively. We denote the number of A individuals by j. Thus N − j represents the number of B individuals. Let us study the evolution of j at one step of the Moran process (for an example, see Figure S1). The transition probabilities associated to the Moran process read [20]:   fAj N − j N j N Πj→j+1 = fAj + fB(N − j) fB(N − j) Πj→j−1 = fAj + fB(N − j) Πj→j = 1 − Πj→j+1 − Πj→j−1 . (S1) The Moran process is a discrete-time Markov process, since the probabilities of states j after one step only depend upon the present value of j. Let us take the limit of continuous time and write the master equation P = RAP giving the probability of being at state j at time t:  −Π0→1 Π0→1 −(Π1→0 + Π1→2) Π1→0 d dt   P0 P1 P2 ... PN   =  0 ... 0 Π1→2 (0) ··· 0 Π2→1 −(Π2→1 + Π2→3) . . . 0 ··· (0) . . . . . . ΠN→N −1 ΠN −1→N −ΠN→N −1 0 ... 0     P0 P1 P2 ... PN   . (S2) This Markov chain has two absorbing states, namely j = 0 and j = N , which correspond to the fixation of B and A individuals, respectively. Once these states are reached, no more changes can occur, in the absence of mutation. It follows that all the components of the first and the last columns of RA equal to 0 (see Eq. S2), so RA is not invertible. In the following, we will denote by eRA the reduced transition rate matrix in which the rows and the columns corresponding to the absorbing states (j = 0, j = N ) are removed, and by eR−1 A its inverse. Let us note that RA is a tridiagonal matrix, which allows for major simplifications of analytical calculations [20]. Note 17 that in order to obtain the transition rate matrix associated to B individuals, one just needs to apply the reversal j ↔ N − j. This corresponds to using the matrix RB = JRAJ where J is the anti-identity matrix. For instance, in 2 dimensions, J =(cid:18)0 1 1 0(cid:19). 2.2 General fixation probabilities and fixation times Definitions. The fixation probability φA j0 represents the probability that A individuals finally succeed and take over the population, starting from j = j0 individuals of type A. In particular, 0 = 0 and φA N = 1. Similarly, φB φA j0 is the fixation probability of the B individuals, still starting from j = j0 individuals of type A. Mean fixation times are the mean times to reach one of the absorbing states. The uncondi- tional fixation time tj0 is the average time until fixation in either j = 0 or j = N , when starting from a number j = j0 of A individuals. The conditional fixation time tA j0 corresponds to the average time until fixation in j = N , when starting from j0, provided that type A fixes. Note that in what follows, we will express the fixation times in numbers of steps of the Moran process. Conversion to generations can then be performed by dividing the number of Moran steps by N . In the following, we present a derivation of the fixation probabilities and of the fixation times in the Moran process [20, 24] that uses the general formalism of mean first passage times [54]. Fixation probabilities. Assuming that at t = 0, the system is at state j = j0, let us focus on the fixation probability φA j0 of the A type in the population. The stochastic process stops at the time bτF P when j fixes, i.e. first reaches one of the absorbing states {j = 0, j = N}. Hence, integrating over all values of bτF P , under the condition that fixation finally occurs in j = N , yields j0 =Z ∞ 0 φA p(bτF P ∈ [t, t + dt] j0 , j∞ = N ) = ΠN −1→NZ ∞ 0 In the last expression, we have taken advantage of the fact that the only way to fix in j = N between t and t + dt is to be in state j = N − 1 at time t and then to transition from N − 1 to N (see Eq. S2). We have thus introduced the probability PN −1(t) of being in state j = N − 1 at time t, starting in state j = j0 at time 0. More generally, the probability Pi(t) can be considered. Integrating the Master equation Eq. S2 to determine Pi(t), with the initial condition Pi(0) = δi j0, where δi j0 denotes the Kronecker delta, which is equal to 1 if i = j0 and 0 otherwise, yields PN −1(t)dt . (S3) j0 of the B type, still starting from j0 (S4) (S5) A )N −1 j0 . φA j0 = −ΠN −1→N (eR−1 A similar reasoning gives the fixation probability φB individuals of type A and N − j0 individuals of type B: j0 = −Π1→0(eR−1 These two probabilities satisfy φA φB j0 + φB A )1 j0 . j0 = 1 since there are 2 absorbing states in the process. Mean fixation times. Let us now focus on the mean fixation times, still assuming that at t = 0, the system is at state j = j0. The probability that fixation in one of the absorbing states {j = 0, j = N} occurs between t and t + dt reads: p(bτF P ∈ [t, t + dt] j0) = N −1Xi=1 Pi(t) − N −1Xi=1 Pi(t + dt) = − N −1Xi=1 dPi dt dt , (S6) where, as above, Pi(t) represents the probability of being in state i at time t starting in j0 at time 0 (note that the initial condition j0 is omitted for brevity). Thus, the unconditional fixation 18 time can be expressed as: tj0 = E[bτF P j0] =Z ∞ 0 Z ∞ 0 N −1Xi=1 t dPi dt = − t p(bτF P ∈ [t, t + dt] j0) N −1Xi=1 Z ∞ Pi(t) dt . 0 dt = (S7) (S8) Here, we used Eq. S6, where the sums run over all the states that are not absorbing (1 ≤ i ≤ N − 1). We also performed an integration by parts, and used [t Pi(t)]∞ 0 = 0 for 1 ≤ i ≤ N − 1, which holds because the probability of reaching an absorbing state of the Markov chain tends Integrating the Master equation Eq. S2 to determine Pi(t), with the initial to 1 as t → ∞. condition Pi(0) = δi j0, gives tj0 = − N −1Xi=1 A )i j0 . (eR−1 p(j j0, j∞ = N ) = φA j φA j0 Pj . (S9) (S11) (S12) To express the conditional fixation time tA j0 of type A, starting from j0 A individuals, we need to take into account the condition that fixation finally occurs in state j = N : p(bτF P ∈ [t, t + dt] j0, j∞ = N ) = The Bayes relation gives: N −1Xi=1 p(i j0, j∞ = N )(t)− N −1Xi=1 p(i j0, j∞ = N )(t + dt) . (S10) By using the same method as for the unconditional fixation time, one obtains: tA j0 = − 1 φA j0 N −1Xi=1 A )i j0 . φA i (eR−1 Similarly, the conditional fixation time of the B type, starting from j0 A individuals, reads: tB j0 = − 1 φB j0 N −1Xi=1 B )i N −j0 . φB N −i(eR−1 It is straightforward to verify that Eqs. S9, S12 and S13 are linked by the relation: tj0 = φB j0tB j0 + φA j0tA j0 . (S13) (S14) Neutral drift. Let us first consider the case without selection fA = fB. In this case, the Moran process can be seen as a non-biased random walk, since individuals of both types are equally likely to be picked for reproduction and death. Fixation eventually happens due to fluctuations. This process, called neutral drift [20] corresponds to diffusion in physics. The transition rates of the system (S1) simplify as follows: Πj→j+1 = Πj→j−1 = j(N − j) Πj→j = 1 − 2 N 2 . j(N − j) N 2 (S15)   Note that here, j can denote the number of A or B individuals indifferently. Indeed, the sym- metry j ↔ N − j entails RA = RB = R, and the transition rate matrix is centrosymmetric, i.e. R = JRJ. For consistency, we will continue to call j the number of A individuals. j0 can be obtained from Eq. S4. It involves elements of the inverse The fixation probability φA of the transition rate matrix. Solving eReR−1 = I, where I is the identity matrix, gives for 1 ≤ i ≤ N − 1 . (S16) (eR−1)N −1 i = − i N N − 1 19 Hence, φA j0 = j0 N . (S17) Taking advantage of the centrosymmetry of R (see above), a property which transfers to eR and eR−1, and entails (eR−1)1 j0 = (eR−1)N −1 N −j0, we can apply Eq. S5, yielding (S18) φB j0 = N − j0 . N Note that φA j0 + φB j0 = 1, as expected. Let us now express the fixation times, focusing on the fate of a single mutant of type B, which corresponds to j0 = N − 1. To compute the unconditional fixation time tN −1, we again need elements of the inverse of the transition rate matrix (see Eq. S9), which are given by Using Eqs. S9 and S19, we obtain: (eR−1)i N −1 = − N N − i . tN −1 = N N −1Xi=1 1 i . (S19) (S20) Similarly, using Eqs. S12, S17 and S19, we obtain the conditional fixation time of type A: tA N −1 = N 2 N − 1 NXi=2 1 i . (S21) Finally, using Eqs. S13, S17 and S19, and making use of the centrosymmetry of eR−1 (see above), yields the conditional fixation time of type B: tB N −1 = N (N − 1) . (S22) Selection. Let us now study the more general case involving selection. For this, let us consider two types A and B having different fitnesses fA and fB, and let us introduce γ = fA/fB. Note that with selection, the transition rate matrices RA and RB = JRAJ are different. In order to compute the fixation probability φA j0, we need some elements of the inverse of the transition A , which are given by: rate matrix eR−1 (eR−1 A )N −1 i = − N N − 1 1 − γ−i 1 − γ−N (cid:0)N − 1 + γ−1(cid:1) for 1 ≤ i ≤ N − 1 . Then, using the previous result and Eq. S4, one obtains: and φA j0 + φB j0 = 1 yields: φA j0 = 1 − γ−j0 1 − γ−N , 1 − γN −j0 1 − γN . elements of the inverse of the transition rate matrix eR−1 φB j0 = Let us now turn to the fixation times. According to Eq. S9, we need to compute other A . Those satisfy: (S23) (S24) (S25) (S26) A )i N −1 = (eR−1 N i(N − i) 1 − γi 1 − γN (i − iγ − N ) for 1 ≤ i ≤ N − 1. 20 Using Eqs. S9 and S26, the unconditional fixation time reads: tN −1 = N 1 − γN N −1Xi=1 (N + iγ − i)(1 − γi) i(N − i) . (S27) To compute the conditional fixation time tA N −1, we substitute Eqs. S24 and S26 in Eq. S12, obtaining: tA N −1 = N (1 − γN )(1 − γ1−N ) N −1Xi=1 (N + iγ − i)(1 − γi)(1 − γ−i) i(N − i) . (S28) A similar reasoning can be used to obtain the conditional fixation time tB N −1 starting from Eq. S13. In order to express the required (eR−1 implies eR−1 1 − γN −i 1 − γN (iγ − i − N γ) for 1 ≤ i ≤ N − 1. B )j 1, we combine the relation eRB = JeRAJ, which A J, together with Eq. S26, and obtain B = JeR−1 (eR−1 (S29) B )i 1 = N i(N − i) This finally yields tB N −1 = N (1 − γN )(1 − γ) N −1Xi=1 (N + iγ − i)(1 − γi)(1 − γN −i) i(N − i) . (S30) 2.3 Fixation probabilities and fixation times used in the main text Let us now make an explicit link between the general expressions obtained above and the fixation probabilities and fixation times used in the main text. Fixation probabilities. First, in the main text, pSR represents the probability that a single resistant (R) mutant fixes without antimicrobial in a population of size N where all other individuals are of type S. Without antimicrobial, fS = 1 and fR = 1 − δ. Considering S as type A and R as type B, we have γ = fS/fR = 1/(1 − δ), and our initial condition is j0 = N − 1. Hence, Eq. S25 yields (S31) (S32) In particular, in the effectively neutral case where δ ≪ 1 and N δ ≪ 1, it yields pSR = φR N −1 = 1 − (1 − δ)−1 1 − (1 − δ)−N . pSR ≈ −δ 1 − e−N log(1−δ) ≈ −δ 1 − eN δ ≈ 1 N , i.e. we recover the result of the neutral case δ = 0 (see Eq. S17). Conversely, in the regime where δ ≪ 1 and N δ ≫ 1, Eq. S31 yields pSR ≈ −δ 1 − eN δ ≈ δe−N δ . (S33) Second, pRC denotes the fixation probability of a single C individual in a population of size N where all other individuals are of type R. Independently of antimicrobial presence, fR = 1− δ and fC = 1. Considering R as type A and C as type B, we have γ = fR/fC = 1 − δ, and our initial condition is j0 = N − 1. Hence, Eq. S25 yields pRC = φC N −1 = δ 1 − (1 − δ)N . In particular, in the effectively neutral case where δ ≪ 1 and N δ ≪ 1, it yields pRC = δ 1 − eN log(1−δ) ≈ δ 1 − e−N δ ≈ 1 N , 21 (S34) (S35) i.e. we again recover the result of the neutral case δ = 0 (see Eq. S17). Conversely, in the regime where δ ≪ 1 and N δ ≫ 1, Eq. S34 yields δ (S36) pRC ≈ 1 − e−N δ ≈ δ . Finally, pSC denotes the fixation probability of a single C mutant in a population of S individuals, without antimicrobial. In this case, fS = fC = 1, so we are in the neutral case, and Eq. S17 yields pSC = 1/N . Fixation times. First, τ d R denotes the average time it takes for the lineage of a single R mutant to disappear in the absence of antimicrobial. Hence, it is equal to the fixation time of the S type in a population that initially contains N − 1 individuals of type S and 1 individual of type R. Considering S as type A and R as type B, we have γ = fS/fR = 1/(1 − δ) without antimicrobial, and our initial condition is j0 = N − 1, so τ d N −1/N (see Eq. S28). Recall that tS N −1 needs to be divided by the population size N because we expressed it in numbers of steps of the Moran process, while τ d R has to be expressed in numbers of generations. While the general formula Eq. S28 is rather complex, in the neutral case δ = 0, it reduces to the much simpler expression in Eq. S21, which yields τ d R ≈ log N for N ≫ 1. For δ > 0, τ d R is shorter than in the neutral case, because the R mutants are out-competed by S individuals. Note that a good approximation to the exact formula in Eq. S28 can be obtained within the diffusion approach [20] (see the Fokker-Planck equation below). R is equal to tS Second, τ f R denotes the average time needed for the R mutants take over with antimicrobial, starting from one R mutant and N − 1 S individuals. Considering S as type A and R as type B, we have γ = fS/fR = 0 with antimicrobial, and our initial condition is j0 = N − 1. Then τ f is equal to tR N −1/N (see Eq. S30), with γ = 0. Using Eq. S30, we obtain R τ f R = N −1Xi=1 1 i , (S37) which entails τ f Finally, τ f R ≈ log N for N ≫ 1. C denotes the average time needed for the C mutants to take over, starting from one C mutant and N − 1 R individuals. Considering R as type A and C as type B, we have γ = fR/fC = 1 − δ, independent whether antimicrobial is present or absent, and our initial condition is j0 = N − 1. Hence, τ f N −1/N (see Eq. S30). In the neutral case δ = 0, tC N −1 reduces to Eq. S22, and thus τ f C ≈ N for N ≫ 1. For δ > 0, it is shorter, as selection favors the fixation of C, and again a good approximation to the exact formula in Eq. S30 can be obtained within the diffusion approach [20] (see the Fokker-Planck equation below). C is given by tC 3 Large populations: deterministic limit If stochastic effects are neglected, the dynamics of a microbial population can be described by coupled differential equations on the numbers of individuals of each genotype [20]. This deterministic approach is appropriate if the number N of competing microorganisms satisfies N µ1 ≫ 1 [34]. Here, we derive and study the deterministic limit of the complete stochastic model studied in the main text. 3.1 From the stochastic model to the deterministic limit Here, we present a full derivation of the deterministic limit of the stochastic model based on the Moran process (see above). This derivation closely follows those of Refs. [23, 24] and is presented here for the sake of pedagogy and completeness. Starting from the Master equation of our stochastic model, we obtain a Fokker-Planck equation, corresponding to the diffusion approximation [20], and then a deterministic differential equation, in the limits of increasingly large population sizes. 22 Let us first recall the Master equation corresponding to the Moran process, where j denotes the number of A individuals and N − j the number of B individuals, as above: dPj(t) dt = Pj−1(t) Πj−1→j + Pj+1(t) Πj+1→j − Pj(t) (Πj→j−1 + Πj→j+1) . (S38) The notations in Eq. S38 are the same as in the previous section, and time is expressed in number of steps of the Moran process. Let us now introduce the reduced variables x = j/N , τ = t/N , as well as ρ(x, τ ) = N Pj(t). Then, since one step of the Moran process occurs each time unit, Eq. S38 can be rewritten as: ρ(x, τ + 1/N ) − ρ(x, τ ) = ρ(x − 1/N, τ ) Π+(x − 1/N ) + ρ(x + 1/N, τ ) Π−(x + 1/N ) − ρ(x, τ )(cid:0)Π−(x) + Π+(x)(cid:1) , (S39) with Π−(x) = Πj→j−1 = fBx(1 − x) fA x + fB (1 − x) and Π+(x) = Πj→j+1 = fAx(1 − x) fA x + fB (1 − x) . (S40) Diffusion approximation. For N ≫ 1, considering that jumps are small at each step of the Moran process, i.e. 1/N ≪ x and 1/N ≪ τ , the probability density ρ(x, τ ) and the transition probabilities Π±(x) can be expanded in a Taylor series around x and τ . This expansion, known as a Kramers-Moyal expansion [55], yields, to first order in 1/N : with ∂ρ(x, τ ) ∂τ ∂ ∂x = − [ρ(x, τ )a(x)] + 1 2 ∂2 ∂x2 (cid:2)ρ(x, τ )b2(x)(cid:3) (S41) a(x) = Π+(x) − Π−(x) and b2(x) = Π+(x) + Π−(x) N . (S42) Eq. S41 is known as a diffusion equation, or a Fokker-Planck equation, or a Kolmogorov forward equation [55], and a(x) corresponds to the selection term (known as the drift term in physics), while b2(x) corresponds to the genetic drift term (known as the diffusion term in physics). Deterministic limit. S41 reduces to: In the limit N → ∞, retaining only the zeroth-order terms in 1/N , Eq. ∂ρ(x, τ ) ∂τ ∂ ∂x = − [ρ(x, τ )a(x)] . (S43) (S44) (S45) (S46) Let us focus on the average value of x, denoted by hxi. Using Eq. S41 yields dhxi dτ 0 ∂τ ∂ρ(x, τ ) =Z 1 = − [x ρ(x, τ ) a(x)]1 = ha(x)i x dx = −Z 1 0 +Z 1 0 0 ∂ ∂x [ρ(x, τ ) a(x)] dx ρ(x, τ ) a(x) dx The first term of right hand side of Eq. S45 vanishes because a(0) = a(1) = 0. In the limit N → ∞, the distribution of x is very peaked around its mean, so hxi ≈ x and ha(x)i ≈ a(x), yielding: dx dτ = x(1 − x) ∆f ¯f , (S47) where ∆f = fA − fB denotes the difference of the fitnesses of the two types, while ¯f = fA x + fB (1 − x) is the average fitness in the population. Eq. S47 is an ordinary differential equation known as the adjusted replicator equation [23]. Recall that τ corresponds to the number t of steps of the Moran process divided by the total number N of individuals in the population. 23 Hence, τ is the time in numbers of generations used in the main text, and Eq. S47 is the proper deterministic limit for our stochastic process. Note that in the framework of the Moran process, fitnesses are only relative. If one wanted to account for absolute fitness effects, so that a whole population reproduces faster if its average fitness is higher, one would need to include an additional rescaling of time τ ′ = τ / ¯f . Note that if ¯f is constant, this rescaling yields a standard replicator equation: dx dτ ′ = x(1 − x)∆f . (S48) 3.2 Deterministic description of the evolution of antimicrobial resistance System of ordinary differential equations. Let us now come back to our model of the evolution of antimicrobial resistance, with three types of microorganisms (see Fig. 1A). In the limit of large populations, the complete stochastic model described in the main text will converge to a deterministic system of ordinary differential equations, as demonstrated above. Generalizing Eq. S48, by considering three types of individuals and taking into account mutations, yields a system of replicator-mutator equations [24]: s = fS(1 − µ1)s − f s r = fR(1 − µ2)r + fS µ1 s − f r s + r + c = 1 , (S49) where s, r and c are the population fractions of S (sensitive), R (resistant) and C (resistant- compensated) microorganisms, respectively, while fS, fR and fC denote their fitnesses, f = fS s + fR r + fC c denotes the average fitness in the population, and dots denote time derivatives. To illustrate that Eq. S49 generalizes Eq. S48, consider the case where c = 0 and µ1 = 0: the first equation of Eq. S48 then yields s = fSs − [fSs + fR(1 − s)]s = s(1 − s)(fS − fR). As demonstrated above, the deterministic limit of our stochastic model yields adjusted replicator equations (see Eq. S47). For the sake of simplicity, the present analytical discussion focuses on standard replicator equations (see Eq. S48). The system of equations Eq. S49 only concerns population fractions, and constitutes the It is mathematically large-population limit N → ∞ of our stochastic model at constant N . convenient to note that the same equations are obtained in the case of a population in which microorganisms have an exponential growth. This model, which enables us to recover the system S49, is governed by the following system of linear differential equations:     NS = fS(1 − µ1)NS NR = fR(1 − µ2)NR + fS µ1 NS NC = fC NC + fR µ2 NR , (S50) where NS, NR and NC are the numbers of sensitive, resistant and resistant-compensated mi- croorganisms, respectively. It is straightforward to show that the population fractions obtained from this exponential growth model satisfy Eq. S49: hence, this simple deterministic model allows one to understand the evolution of large microbial populations described by the Moran model (even though the total population is constant in the Moran model). Analytical resolution. Being linear, the system in Eq. S50 is straightforward to solve ana- lytically:   S R C  =   0 0 1 0 1 fR µ2 1 fS µ1 fS(1−µ1)−fR(1−µ2) fS µ1 fR µ2 fR(1−µ2)−fC (fS(1−µ1)−fR(1−µ2))(fS(1−µ1)−fC)     β1 efC t β2 efR(1−µ2)t β3 efS(1−µ1)t   (S51) where β1, β2 and β3 can be expressed from the initial conditions S(0), R(0) and C(0). The fractions s, r and c can then be obtained from this solution, e.g. through s = S/(S + R + C). 24 Limiting regimes and characteristic timescales. As in the main text, we are going to focus on the case where the population initially only comprises sensitive microorganisms, i.e. s(0) = 1. In the case of periodic alternations of absence and presence of antimicrobial, a small fraction of R microorganisms will appear within the first half-period without antimicrobial. The subsequent evolution of the population composition can be separated into three successive regimes. In the first one, it suffices to consider S and R microorganisms, as the fraction of C is negligible, because the appearance of C requires an additional mutation. The second regime is more complex, and involves all three types of microorganisms, as the growth of C microorganisms makes the fractions of S and R microorganisms decrease. Then, provided that antimicrobial has been present for a sufficient time, the fraction of S microorganisms becomes negligible, because they cannot divide with antimicrobial. Hence, the third regime only involves R and C microorganisms, and does not depend on the presence or absence of antimicrobial, because the fitnesses of R and C are unaffected. Here, we determine analytically the main timescales involved in these first and third regimes. First regime: S vs. R. Let us consider the first regime where there are almost only S and R microorganisms. We are interested in the population fractions s(t) and r(t), with s(t)+ r(t) ≈ 1. Eq. S49 then gives: (S52) where we have defined ∆f1 = fS(1 − µ1) − fR and ∆f2 = fS − fR. Note that we expect ∆f1 ≈ ∆f2, since biologically relevant values generally satisfy µ1 ≪ 1 and µ1 ≪ δ. The solution of Eq. S52 reads s = s (∆f1 − s ∆f2) , s(t) = 1 − s0 s0 e∆f1t + s0 ∆f2 ∆f1 ∆f2 ∆f1 , e∆f1t (S53) where s0 is the fraction of S microorganisms at the beginning of the first regime (taken as t = 0 here). In the presence of antimicrobial (fS = 0), the previous expression can be simplified, using ∆f1 = ∆f2 = −(1 − δ). This allows us to identify the characteristic time τ1 of the decay of s, as R microorganisms take over: The duration t1 of the first regime in the presence of antimicrobial is governed by τ1. More τ1 = −1 ∆f1 = 1 1 − δ . (S54) precisely, Eq. S53 yields: t1 = 1 1 − δ log(cid:18)s0 (1 − s1) s1 (1 − s0)(cid:19) , (S55) where s1 is the fraction of S microorganisms at the end of the first regime, at which point the fraction of C microorganisms is no longer negligible. Third regime: R vs. C. Let us now turn to the third regime, assuming that antimicrobial has been present for a long enough time to allow S microorganisms to become a small minority. Eq. S49 then gives: (S56) with ∆f3 = fR(1 − µ2) − fC = −δ(1 − µ2) − µ2 and ∆f4 = fR − fC = −δ, independently of whether antimicrobial is present or not. Again, we generally expect ∆f3 ≈ ∆f4. The solution of Eq. S56 reads r = r (∆f3 − r ∆f4) , r(t) = r2 (1 − µ2 + µ2/δ) e−(δ(1−µ2)+µ2)t 1 − µ2 + µ2/δ − r2 + r2 e−(δ(1−µ2)+µ2)t ≈µ2≪1 µ2≪δ r2 e−δt 1 − r2 + r2 e−δt , (S57) 25 where r2 is the fraction of R microorganisms at the beginning of the third regime (taken as t = 0 here). Hence, the characteristic time τ3 of the decay of r reads: τ3 = 1 µ2 + δ(1 − µ2) ≈µ2≪1 µ2≪δ 1 δ . (S58) The duration t3 of the third regime in the presence of antimicrobial is governed by τ3. More precisely, Eq. S57 yields: 1 δ t3 ≈µ2≪1 µ2≪δ log(cid:18) r2 (1 − r3) r3 (1 − r2)(cid:19) . (S59) where r3 is the fraction of R microorganisms at the end of this regime, when C has become dominant in the population. Note that the timescales obtained here are governed by selection (through the relevant fitness differences δ and 1 − δ). This stands in contrast with the results from our stochastic model (see main text) where mutation rates are crucial, especially through the waiting time before resistant mutants appear. In the deterministic description considered here, small fractions of resistant mutants appear right away, so this consideration is irrelevant. However, mutation rates come into play in the durations of the different regimes within the deterministic model, through the fractions of each type of microorganisms at the beginning and at the end of each regime, but with a weak logarithmic dependence (see Eqs. S55-S59). 3.3 Comparison of stochastic and deterministic results As in the main text, we now focus on the impact of a periodic presence of antimicrobial on the time it takes for a population to fully evolve resistance. For large microbial populations satisfying N ≫ 1/µ1, we wish to check that the system of differential equations in Eq. S49 recovers the results obtained with our stochastic model. To this end, we solve the system in Eq. S49 numerically in the case of a periodic presence of antimicrobial. Note that complete fixation of a genotype does not happen in the deterministic model. Conversely, in the stochastic model, for a population of size N , the fixation of C corresponds to the discrete Moran step where the fraction c jumps from 1 − 1/N to 1. Hence, for our comparison between the deterministic results and the stochastic ones obtained for N microorganisms, we consider that C effectively fixes in the deterministic model when the fraction c reaches 1− 1/N . In addition, for exactness, we use a numerical resolution of the system in Eq. S49 where time is rescaled through t → t/ ¯f . Indeed, the proper deterministic limit of our stochastic model corresponds to modified replicator equations, such as Eq. S47 (see above). Let us now present an analytical approximation for tf Fig. S2 shows that the deterministic model yields results very close to those obtained through the stochastic model, in the case of large population sizes N ≥ 1/µ1. We recover the regimes described in the main text, with a plateau for short periods, and a linear dependence on T for larger ones. Moreover, the relative error made by using the deterministic model instead of the stochastic one is less than ∼ 20% (resp. ∼ 10%) for all data points with N = 105 (resp. N = 106) in Fig. S2. C, based on the different timescales computed previously. As the population is initially only composed of S microorganisms, they will remain dominant during the first half-period without antimicrobial, since they are fitter than R mutants (and we assume that T /2 is not large enough to extend to the point where C starts being important, which would then correspond to the valley crossing case). Afterwards, R microorganisms start growing fast during the second half-period. Note that in the deterministic case, there is always a nonzero fraction of resistant microorganisms at the end of the first half- period without antimicrobial, contrary to the stochastic case studied in the main text. Hence, we compute the fraction s0 = s(T /2) of S microorganisms at the end of the first half period, by using results for the above-described first regime without antimicrobials. This fraction s0 = s(T /2) is then taken as the initial condition of the first regime with antimicrobial. Then, for simplicity, 26 Figure S2: Large populations: stochastic model vs. deterministic model. The total time tf C of full resistance evolution is plotted versus the period T of alternations of absence and presence of antimicrobial, in the case of symmetric alternations. Results from simulations of the stochastic model (see Fig. 3A), numerical resolution of the deterministic model, and an analytical approximation of the deterministic solution (Eqs. S60-S61), are represented for N = 105 (A) and N = 106 (B). Parameter values: µ1 = 10−5, µ2 = 10−3, and δ = 0.1. we assume that s decays until it reaches s1 ≈ 0.1 (so r1 ≈ 0.9), while remaining in the first regime described above, in the presence of antimicrobial. We then assume the duration of the second regime is negligible, and consider that the third regime process starts right away, with a fraction r2 ≈ 0.9. As explained above, we consider that the third regime ends upon effective fixation of C, i.e. when c reaches 1 − 1/N , which implies r3 = 1/N . Using Eqs. S55 and S59, we obtain: (S60) (S61) log (9(N − 1)) , where s(T /2) is obtained by using Eq. S53 in the absence of antimicrobial: + log(cid:18) 9 s(T /2) 1 − s(T /2)(cid:19) + 1 δ tf C ≈ T 2 1 1 − δ s(T /2) = e(µ2+δ(1−µ2)−µ1)T /2 1 − µ2+δ(1−µ2) µ2+δ(1−µ2)−µ1 (cid:0)1 − e(µ2+δ(1−µ2)−µ1)T /2(cid:1) . Eqs. S60-S61 yield good approximations of the analytical results obtained by numerical resolution of Eq. S49, as can be seen on Fig. S2. More precisely, the relative error made by using this approximation instead of the full numerical resolution is less than ∼ 13% for all parameters in Fig. S2. For T ≫ 2/δ, Eq. S61 reduces to s(T /2) ≈ 1− µ1/[µ2 + δ(1− µ2)] ≈ 1− µ1/δ, so only the first term in Eq. S60 then depends on T . Hence, this term becomes dominant for large T , yielding tf C ≈ T /2 in this limit. This asymptotic behavior is again consistent with our predictions from the stochastic model (see main text). The horizontal purple solid line at large T in Fig. 3A, and the horizontal solid lines at large N in Fig. 3B, both correspond to tf C ≈ T /2, showing excellent agreement with our stochastic simulations as well. Conversely, for small periods, the first term of Eq. S60 can be neglected, so the dependence on T of tf C is weaker (Eq. S61 reduces to s(T /2) ≈ 1− µ1T /2 for T ≪ 2/δ, so a weak logarithmic dependence on T remains, due to the second term of Eq. S60). It is interesting to note that the third term of tf C in Eq. S60 also increases logarithmically with N . This stands in contrast with the case of smaller populations, where our stochastic study showed that tf C essentially decreases linearly with N (see main text). This change of behavior as N increases can be seen on Fig. 3A in the regime of small T (in particular, for large N , the purple data points corresponding to N = 106 are then slightly higher than the blue ones corresponding to N = 105; see also Fig. S2, where the y-axis range and scale are the same on panels A and B). 27 4 Comparison to spontaneous fitness valley crossing 4.1 No antimicrobial: Crossing of a symmetric fitness valley Let us compare the alternation-driven evolution of resistance to what would happen in the absence of alternations of phases of absence and presence of antimicrobial. If a population composed only of S (sensitive) microorganisms is subjected to a continuous presence of antimi- crobial, it will not evolve resistance, because divisions are blocked (see Fig. 1A). Conversely, a population of S microorganisms that is never subjected to antimicrobial can spontaneously evolve resistance. In our model, this will eventually happen. This process is difficult and slow, because of the initial fitness cost of resistance: it requires crossing a fitness valley (see Fig. 1A). Fitness valley crossing has been studied in detail [22, 42, 56 -- 58], but usually in the case where the final mutant has a higher fitness than the initial organism. In the evolution of antimicro- bial resistance, compensatory mutations generally yield microorganisms with antimicrobial-free fitnesses that are similar to, but not higher than those of sensitive microorganisms [3, 4, 6]. Hence, we here extend the known results for fitness valley crossing by constant-size homoge- neous asexual populations [22] to "symmetric" fitness valleys, where the final genotype has no selective advantage compared to the initial one. Briefly, the main difference with Ref. [22] is that the probability of establishment of the second mutant (C) in a population with a majority of non-mutants (S) is 1/N instead of being given by the selective advantage s of the second mutant. This probability plays an important role in the tunneling case. There are two different ways of crossing a fitness valley. In sequential fixation, the first deleterious mutant fixes in the population, and then the second mutant fixes. In tunneling [56], the first mutant never fixes in the population, but a lineage of second mutants arises from a minority of first mutants, and fixes. For a given valley, characterized by δ (see Fig. 1A), popu- lation size N determines which mechanism dominates. Sequential fixation requires the fixation of a deleterious mutant through genetic drift, and dominates for small N , when stochasticity is important. Tunneling dominates above a certain N [22, 57]. Let us study these two mechanisms in the regime of rare mutations N µ1 ≪ 1 where stochasticity is crucial. In sequential fixation, the average time τSF to cross a valley is the sum of those of each step involved [22]. Hence τSF = 1/(N µ1pSR) + 1/(N µ2pRC), where pSR (resp. pRC) is the fixation probability of a single R (resp. C) individual in a population of size N where all other individuals are of type S (resp. R). Fixation probabilities are known in the Moran process (see Supplementary Material, Section 2). In particular, if N δ ≪ 1 then pSR ≈ pRC ≈ 1/N for our symmetric valley, so τSF ≈ 1/µ1 + 1/µ2 (≈ 1/µ1 if µ1 ≪ µ2), while if δ ≪ 1 and N δ ≫ 1 then pSR ≈ δe−N δ and pRC ≈ δ ≫ pSR, so τSF ≈ eN δ/(N µ1δ). In tunneling, the key timescale is that of the appearance of a successful first (R) mutant, i.e. a first mutant whose lineage will give rise to a second (C) mutant that will fix in the population [22]. Neglecting subsequent second mutation appearance and fixation times, the average tunneling time reads τT ≈ 1/(N µ1p1), where p1 is the probability that a first mutant is successful [22]. Upon each division of a first mutant, the probability of giving rise to a second mutant that will fix is p = µ2pSC, where pSC is the fixation probability of a single C mutant in a population of S individuals. For our symmetric valley, pSC = 1/N , so p = µ2/N . In the neutral rate at which successful first mutants are produced. Since the lineage of each new first mutant case δ = 0, Ref. [22] demonstrated that the first-mutant lineages that survive for at least ∼ 1/√p generations, and reach a size ∼ 1/√p, are very likely to be successful, and fully determine the has a probability ∼ √p of surviving for at least ∼ 1/√p generations [22], the probability that a first mutant is successful is p1 ∼ √p ∼pµ2/N . If δ > 0, a first mutant remains effectively neutral if its lineage size is smaller than 1/δ [22]. Hence, if δ < pµ2/N , p1 ∼pµ2/N still holds. (This requires N µ2 ≫ 1, otherwise the first mutant fixes before its lineage reaches a size pN/µ2.) Finally, if δ > pµ2/N , the lineage of a first mutant will reach a size at most ∼ 1/δ, with a probability ∼ δ and a lifetime ∼ 1/δ [22], yielding p1 ∼ µ2/(N δ). Given the substantial cost of resistance mutations (δ ∼ 0.1 [4, 6]) and the low compensatory mutation rates (in bacteria µ2 ∼ 10−8 [4]), let us henceforth focus on the case where δ >pµ2/N 28 (which is appropriate for all N ≥ 1 with the values mentioned). Then τT ≈ 1/(N µ1p1) ≈ δ/(µ1µ2), and two extreme cases can be distinguished: (A) N δ ≪ 1 (effectively neutral regime): Then, τSF ≈ 1/µ1 (for µ1 ≪ µ2) and τT ≈ δ/(µ1µ2). Given the orders of magnitude above, generally δ > µ2 in resistance evolution. Hence, sequential fixation is fastest, and the valley crossing time τV reads: τV = τSF ≈ 1 µ1 . (B) δ ≪ 1 and N δ ≫ 1: Then, τV = min (τSF, τT) ≈ min(cid:18) eN δ N µ1δ (S62) (S63) δ µ1µ2(cid:19) . , The transition from sequential fixation to tunneling [22] occurs when N δe−N δ = µ2/δ. We have focused on the rare mutation regime N µ1 ≪ 1. If mutations are more frequent, the first successful lineage of R mutants that appears may not be the one that eventually fixes, so the valley-crossing time becomes shorter [22]. In Fig. 3B, the black simulation data points were obtained without any antimicrobial. The population then evolves resistance by valley crossing. The black curves correspond to our ana- lytical predictions in Eq. S62 for N ≪ 1/δ and in Eq. S63 for N ≫ 1/δ. In the latter regime, the transition from sequential fixation to tunneling occurs at N ≈ 65 for the parameters of Fig. 3B. The agreement between simulation results and analytical predictions is excellent, with no adjustable parameter. 4.2 Alternation-driven process vs. valley-crossing process Now that we have studied the spontaneous crossing of a symmetric fitness valley without any antimicrobial, let us come back to our periodic alternations of phases of absence and presence of antimicrobial. Resistance can then evolve by two distinct mechanisms, namely the alternation- driven process and the spontaneous valley-crossing process. It is important to compare the associated timescales, in order to assess which process will happen faster and dominate. This will shed light on the acceleration of resistance evolution by the alternations. For generality, we consider asymmetric alternations. With alternations, spontaneous valley crossing can still happen, but new R lineages cannot appear with antimicrobial, because S individuals cannot divide (see Fig. 1A). Since the appear- ance of a successful R mutant is usually the longest step of valley crossing (see above), the average valley crossing time τ ′ V with alternations will be longer by a factor T /T1 than that with- out antimicrobial (τV), if more than one antimicrobial-free phase is needed to cross the valley, i.e. if T1 ≪ τV. Eqs. S62 and S63 then yield τ ′ V ≈ τ ′ V ≈ T T1µ1 T T1 min(cid:18) eN δ N µ1δ δ µ1µ2(cid:19) , for N δ ≪ 1 , for δ ≪ 1 and N δ ≫ 1 . (S64) (S65) Conversely, if T1 ≫ τV, valley crossing generally happens within the first antimicrobial-free phase. Hence, the average valley crossing time τV is given by Eqs. S62 and S63. (Recall that the process is assumed to begin with an antimicrobial-free phase.) R it takes to first observe an R organism in the presence of antimicrobial, i.e. tf We can now compare the timescales of the valley-crossing process to those of the alternation- driven process. For simplicity, let us assume that the dominant timescale in the latter process is the time ta C ≈ ta (see Eq. 2). This is the case in a large and relevant range of parameters, especially if µ1 ≪ µ2, as discussed above. Note also that the final step of fixation of the successful C lineage, which can become long in large populations (up to ∼ N in the neutral case, see Supplementary Material, Section 2), is the same in the alternation-driven process and in the valley-crossing process, so it R 29 R (recall that τ d R/τ ′ R/τ ′ Indeed, if N δ ≪ 1, τ ′ V is given by Eq. S64, so for all N > 1, ta V for T1 ≪ τV, where Eqs. S64 and S65 hold. does not enter the comparison. The expression of ta R in Eq. 3 should thus be compared to the valley crossing time. If T1 ≫ τV, valley crossing happens before any alternation, and is thus the relevant process, with time τV given by Eqs. S62 and S63. Let us now conduct our comparison of ta V ≈ µ2/(N δ) ≪ 1 in the tunneling regime. Hence, if T1 ≪ τ d R and τ ′ (A) If T1 ≪ τ d R is the average lifetime of an R lineage without antimicrobial, before it goes extinct): The alternation-driven process, with timescale ta R = T /(N µ1T1) (see R < τ ′ V. Eq. 3), dominates. And if N δ ≫ 1 and δ ≪ 1, Eq. S65 yields ta V ≈ δe−N δ ≪ 1 in the sequential fixation regime, and ta R, the alternation- driven process dominates. Thus, alternations of absence and presence of antimicrobial strongly accelerate resistance evolution. For instance, in Fig. 3A, for N = 100 and T /2 ≪ τ d R, the alternation-driven process takes ta R = 2/(N µ1) = 2 × 103 generations, while valley crossing takes τV = δ/(µ1µ2) = 107 generations without antimicrobial: alternations yield a speedup of 4 orders of magnitude. The speedup is even stronger for larger populations. Conversely, for T1 ≪ T2, while the alternation-driven process is shorter than the valley-crossing process in the presence of alternations, it can nevertheless be longer than the valley-crossing process in the absence of antimicrobial. In this case, the drug actually slows down the evolution of resistance. R, in the tunneling regime, provided that 1/N ≪ δ ≪ 2pµ2/N , When T1 ≪ T2 and T1 ≪ τ d valley crossing takes δ/(µ1µ2) in the absence of antimicrobial (see Eq. S63), and T2δ/(T1µ1µ2) in the presence of alternations satisfying T1 ≪ T2 (see Eq. S65). Meanwhile, the switch-driven process takes T2/(T1N µ1) (see above). Hence, if T2 ≫ T1N δ/µ2, the alternation-driven process dominates, but it is slower than the valley-crossing process in the absence of antimicrobial: the drug then slows down resistance evolution. This effect can be seen on Fig. 4 for T1 ≪ T2 and T1 ≪ τ d R. R) (see Eq. 3). If N δ ≪ 1, valley crossing by sequential fixation is the dominant process. Indeed, Eq. S64 yields ta R ≫ 1. If N δ ≫ 1 and δ ≪ 1, Eq. S65 yields ta V ≈ µ2T1/(N δτ d R) in the tunneling regime. A transition from the alternation-driven process to valley crossing occurs when these ratios reach 1. Qualitatively, if N is large enough and/or if T1 is short enough, the alternation-driven process dominates. R in the sequential fixation regime, and ta R/τ ′ V ≈ δe−N δT1/τ d R/τ ′ V ≈ T1/τ d (B) If T1 ≫ τ d R: Then ta R = T /(N µ1τ d R/τ ′ R/τ ′ V reaches 1 for T1 = N δτ d R, extinction events occur when T1 ≫ τ d For example, in Fig. 4A, parameters are such that the dominant mechanism of valley crossing R/µ2 ≈ 2.6 × 105 generations. This transition to is tunneling, so ta the valley-crossing plateau is indeed observed for the curves with large enough T2. (Recall that if T2 ≪ τ f R, see Fig. S3B.) The black horizontal lines in Figs. 4A and 4B correspond to our analytical prediction in Eq. S65, giving τ ′ V ≈ δ/(µ1µ2) if T1 ≫ max(T2, τ d R). Similarly, in Fig. 3A, horizontal solid lines at large T correspond to the valley crossing times in Eqs. S64 or S65, depending on N . In Fig. 3B, in the regime of small N and large T , resistance evolution is achieved by tunneling-type valley crossing, yielding a plateau in the neutral regime N ≪ 1/δ (see Eq. S64, plotted as a horizontal purple line) and an exponential increase for intermediate N (see Eq. S65). For larger N , we observe a T -dependent transition to the alternation-driven process, which can be fully understood using the ratio ta R/τ ′ V (see above). 5 Detailed analysis of asymmetric alternations 5.1 Particular regimes Here, we examine whether R mutants will fix during a single phase with antimicrobial, of duration T2. The fixation time of the lineage of an R mutant in the presence of antimicrobial is τ f R, fixation will happen within T2. In the opposite case, the fixation of R is not likely to occur within a single phase with antimicrobial. Two situations exist in this case (see Fig. S3). R ≈ log N for N ≫ 1 [20] (see above). If T2 ≫ τ f (A) If T2 ≪ τ f R (Fig. S3A): The R lineage will drift for multiple periods, but its extinction is unlikely, as for symmetric alternations. This effect can induce a slight increase R and T1 ≪ τ d 30 Figure S3: Particular regimes. The number of R individuals in the population is plotted versus time under alternations of phases without (white) and with antimicrobial (gray). Data extracted from simulation runs. (A) T2 ≪ τ f R: the R lineage drifts for multiple periods. Parameters: N = 103, T1 = 10−1, T2 = 10−2 . (B) T2 ≪ τ f R: the R lineage goes extinct. Parameters: N = 102, T1 = 102, T2 = 1. In both (A) and (B), µ1 = 10−5, µ2 = 10−3 and δ = 0.1. R and T1 ≫ τ d R and T1 ≪ τ d of the total time of resistance evolution, which is usually negligible. (B) If T2 ≪ τ f R and T1 ≫ τ d R (Fig. S3B): The R lineage is likely to go extinct even after it has started growing in the presence of antimicrobial. This typically implies T1 ≫ T2, since τ f R ≈ log N and τ d R . log N for N ≫ 1 (see above). Hence, this case is specific to (very) asymmetric alternations. Spontaneous valley crossing then becomes the fastest process of resistance evolution (see Supplementary Material, Section 4). 5.2 Varying T2 at fixed T1 In the main text, we present a detailed analysis of what happens when T1 is varied at fixed T2 (see Fig. 4A). Here, we present a similar analysis if T2 is varied at fixed T1. Fig. 4B shows the corresponding simulation results, together with our analytical predictions from Eqs. 2 and 3. In particular, a minimum is observed in Fig. 4B when varying T2 for T1 ≫ τ d R: R ≪ T1, valley crossing dominates. • When T2 ≪ τ d • When τ d from T2. R ≪ T2 ≪ T1, Eq. 3 gives ta R = T /(N µ1τ d R) ≈ T1/(N µ1τ d R), which is independent • As T2 is further increased, ta R = T /(N µ1τ d R) increases, becoming proportional to T2 when T2 ≫ T1. R is T1/(N µ1τ d R) and is attained for τ d Hence, the minimum of ta case where T1 ≪ τ d for T2 ≪ T1 ≪ τ d (as seen above). Then, ta always slower than the alternation-driven process when T1 ≪ τ d expected at large T2 in this case. R ≪ T2 ≪ T1. In the opposite R, Eq. 3 still gives ta R = 1/(N µ1) R, which means that the first R mutant yields the full evolution of resistance R becomes proportional to T2 for T2 ≫ T1. Note that valley crossing is R (see above), so no plateau is R = T /(N µ1T1). Thus, ta R reaches a plateau ta 6 Robustness of the binary antimicrobial action model Throughout our study, we have modeled the action of the antimicrobial in a binary way: below the MIC ("absence of antimicrobial"), growth is not affected, while above it ("presence of an- timicrobial"), sensitive microorganisms cannot grow at all (see Model section in the main text). The relationship between antimicrobial concentration and microorganism fitness is termed the pharmacodynamics of the antimicrobial [16, 49]. Our binary approximation is motivated by the usual steepness of pharmacodynamic curves around the MIC [16]. However, this steepness is 31 not infinite, and it is different for each antimicrobial. Here, we investigate the robustness of our binary model. If one goes beyond the binary model and accounts for the smoothness of the pharmacody- namic curve, one additional factor enters the determination of the time dependence of fitness. It is the time dependence of the antimicrobial concentration, typically in a treated patient, which is known as pharmacokinetics [16, 49]. In fact, the time dependence of the fitness of sensitive mi- croorganisms will be determined by a combination of pharmacodynamics and pharmacokinetics. Experimental pharmacodynamic curves are well-fitted by Hill functions, and pharmacokinetic curves are often modeled by exponential decays of drug concentration after intake [16]. The fitness versus time curve upon periodic antimicrobial intake will be a smooth periodic function resulting from the mathematical function composition of these two empirical relationships. The main feature of this curve will be how smooth or steep it is, which can be characterized by its rise time, i.e. the time it takes to rise from a value of fS close to 0 to one close to 1. Recall that the fitness fS of sensitive microorganisms ranges between 0 at very high antimicrobial concen- trations and 1 without antimicrobial. In practice, we chose to define the rise time as the time taken to rise from fS = 0.1 to fS = 0.9. Thus motivated, we consider a smooth and periodic fitness versus time relationship fS(t) (see Fig. S4A), and we study the impact of the rise time Θ on the evolution of antimicrobial resistance in a microbial population. In practice, our smooth function, shown in Fig. S4A, is du, such that over each period of duration 0 e−u2 T : built using the error function erf(x) = (2/√π)R x 2(cid:19)(cid:19)(cid:21) fS(t) = 1 − (t − nT − T )(cid:19)(cid:21) 2(cid:20)1 + erf(cid:18) 2 2(cid:20)1 + erf(cid:18) 2 Θ(cid:18)t − nT − fS(t) = 1 1 Θ T if nT + if nT + , 3T 4 T 4 ≤ t < nT + 3T 4 ≤ t < (n + 1)T + (S66) (S67) T 4 , where n is a non-negative integer. In addition, we take fS(t) = 1 for 0 ≤ t ≤ T /4, i.e. we start without antimicrobial at t = 0, and the first decrease of fitness occurs around t = T /2, in order to be as close as possible to our binary approximation (see Fig. 1B). Finally, as an extremely smooth case, we consider the case of a fitness fS modeled by a sine function of period T , with the same initial condition and phase as our function with variable smoothness. Figure S4: Robustness of the binary antimicrobial action model. (A) Smooth and periodic fitness versus time relationship considered: Θ denotes the rise time. (B) Total time tf C of full resistance evolution versus the period T for smooth alternations with different values of Θ, and for the binary model. Data points correspond to the average of simulation results (over 10 to 103 replicates), and error bars (often smaller than markers) represent 95% confidence intervals. Parameter values: µ1 = 10−5, µ2 = 10−3, δ = 0.1, and N = 100. We have performed stochastic simulations using the model described in the main text, but with the fitness versus time relationship given in Eqs. S66-S67. Fig. S4 shows that for small 32 rise times Θ, the dependence on the period T of the total time tf C of full resistance evolution is the same as with our binary approximation, provided that the rise time is much smaller than the period, Θ ≪ T . Conversely, for small Θ satisfying Θ ≥ T , in which case our function is very smooth even though the absolute rise time is short, the behavior of tf C is similar to that obtained for the sine function. For larger values of Θ, namely Θ ≫ 10, the binary case is no longer matched when Θ ≪ T , and instead, a behavior intermediate between the binary case and the sine case is observed. This intermediate behavior gets closer to that observed in the sine case as Θ is increased. R, τ f R and 1/(N µ1), the shortest ones being τ d These results can be rationalized as follows. When Θ is smaller than the relevant evolutionary timescales identified in the main text (τ d R and τ f R for N µ1 ≪ 1), no relevant evolutionary process process can happen during a single smooth rise or decay of the fitness. If in addition Θ is much smaller than the environmental timescale T , then the fitness versus time function is steep and effectively binary. However, if Θ is not much smaller than T , then the function is smooth, and the binary approximation is inappropriate. Finally, if Θ is longer than the shortest relevant evolutionary timescales (τ d R), then relevant evolutionary processes can happen within a single smooth rise or decay of the fitness, and the behavior is more complex. In a nutshell, our binary approximation is appropriate provided that the rise time satisfies Θ ≪ min(T, τ d R, τ f R, τ f R). 33
1904.04033
1
1904
2019-03-25T08:38:02
Does electronic coherence enhance anticorrelated pigment vibrations under realistic conditions?
[ "physics.bio-ph", "quant-ph" ]
The light-harvesting efficiency of a photoactive molecular complex is largely determined by the properties of its electronic quantum states. Those, in turn, are influenced by molecular vibrational states of the nuclear degrees of freedom. Here, we reexamine two recently formulated concepts that a coherent vibronic coupling between molecular states would either extend the electronic coherence lifetime or enhance the amplitude of the anticorrelated vibrational mode at longer times. For this, we study a vibronically coupled dimer and calculate the nonlinear two-dimensional (2D) electronic spectra which directly reveal electronic coherence. The timescale of electronic coherence is initially extracted by measuring the anti-diagonal bandwidth of the central peak in the 2D spectrum at zero waiting time. Based on the residual analysis, we identify small-amplitude long-lived oscillations in the cross-peaks, which, however, are solely due to groundstate vibrational coherence, regardless of having resonant or off-resonant conditions. Our studies neither show an enhancement of the electronic quantum coherence nor an enhancement of the anticorrelated vibrational mode by the vibronic coupling under ambient conditions.
physics.bio-ph
physics
Does electronic coherence enhance anticorrelated pig- ment vibrations under realistic conditions? Hong-Guang Duan1,2,3, Michael Thorwart2,3 & R. J. Dwayne Miller1,3,4 1Max Planck Institute for the Structure and Dynamics of Matter, Luruper Chaussee 149, 22761, Hamburg, Germany 2I. Institut fur Theoretische Physik, Universitat Hamburg, Jungiusstrasse 9, 20355 Hamburg, Ger- many 3The Hamburg Center for Ultrafast Imaging, Luruper Chaussee 149, 22761 Hamburg, Germany 4The Departments of Chemistry and Physics, University of Toronto, 80 St. George Street, Toronto Canada M5S 3H6 April 9, 2019 The light-harvesting efficiency of a photoactive molecular complex is largely determined by the properties of its electronic quantum states. Those, in turn, are influenced by molecu- lar vibrational states of the nuclear degrees of freedom. Here, we reexamine two recently formulated concepts that a coherent vibronic coupling between molecular states would ei- ther extend the electronic coherence lifetime or enhance the amplitude of the anticorrelated vibrational mode at longer times. For this, we study a vibronically coupled dimer and cal- culate the nonlinear two-dimensional (2D) electronic spectra which directly reveal electronic coherence. The timescale of electronic coherence is initially extracted by measuring the anti- diagonal bandwidth of the central peak in the 2D spectrum at zero waiting time. Based on 1 the residual analysis, we identify small-amplitude long-lived oscillations in the cross-peaks, which, however, are solely due to groundstate vibrational coherence, regardless of having resonant or off-resonant conditions. Our studies neither show an enhancement of the elec- tronic quantum coherence nor an enhancement of the anticorrelated vibrational mode by the vibronic coupling under ambient conditions. In the initial steps of photosynthesis, photoactive molecular complexes capture the sunlight energy and transfer it to the reaction center on an ultrafast time scale and with unity quantum ef- ficiency 1. The performance is determined by the molecular electronic properties, in concert with the molecular vibrations and coupling to the environment given by a solvent and the surrounding pigments and proteins. To investigate the energy transfer, ultrafast 2D electronic spectroscopy 2 -- 4 is able to resolve fs time scales. It is able to reveal the interactions between the energetically close- by lying molecular electronic states, for which the linear spectra are commonly highly congested and broadened by the strong static disorder 5. Recent experimental studies of the Fenna-Matthews- Olson (FMO) complex reported long-lived oscillations of the cross-peaks both at low 6 and at room temperature 7 which have been assigned to enhanced electronic coherence. This has gen- erated tremendous interest in this new field of quantum biology 8, aiming to reveal a functional connection between photosynthetic energy transfer and long-lived quantum coherence. Moreover, also in photoactive marine cryptophyte algae 9, the light-harvesting complex LHCII 10 and in the Photosystem II reaction center 11, 12, long-lived oscillations have been experimentally reported at low and room temperature. 2 To model the reported 6 coherence, Ishizaki and Fleming have used a parametrized model of the FMO complex 13, with a rather small reorganization energy of 35 cm−1 to fit the electronic coherence timescale 14. This value was extracted 15 from flourescence line narrowing measure- ments at low temperature 16 and does not include high-frequency intramolecular modes. However, even with the small reorganization energy, Shi et al. have calculated the complete 2D spectra and found a much shorter electronic coherence lifetime 17. They have pointed out that the interpretation of the long-lived coherence could just be due to the intentional magnification of the 2D spectral amplitudes by the deliberately used inverse hyperbolic sine scale. In addition, electronic quantum coherence has been questioned to play any crucial role for the energy transfer as the transport is domainted by largely incoherent exciton relaxation 18, 19. A critical issue has been the use of an inadequate spectral distribution of the environmental fluctuations. The experimentally determined spectral density with a larger reorganization energy 15 has been used to calculate the dynamics by the quasiadiabatic propagator path integral 18. There, a local vibrational mode at 180 cm−1 with a broadening of 29 cm−1 has been included, with a total reorganization energy of 100 cm−1. The numerically exact results also show a significantly shorter electronic coherence lifetime. Re- cent QM/MM-simulations 20 -- 24 yield site-resolved spectral densities with reorganization energies of 150 to 200 cm−1. Thus, theoretical studies showed that pure electronic quantum coherence can not survive under ambient conditions. Motivated by this disagreement, the coherent exciton dynamics in the FMO complex has been reexamined experimentally by 2D electronic spectroscopy 25. A fit to an Ohmic spectral density with a broadened high-frequency mode yields a reorganization energy of 3 190 cm−1. The observed lifetime of the electronic coherence of ∼60 fs is too short to play any functional role in the energy transport, which occurs of the ps time scale. In addition to the electronic coherence, signatures of the vibrational coherence of the pigment- protein host can also be accessed on the same spectroscopic footing 26 -- 29. Yet, electronic coherence can be distinguished from vibrational coherence 30, 33, 34. Long-lived pure electronic coherence is unexpected to exist in most light harvesting complexes. However, long-lived vibrational coherence is common and is not expected to strongly affect light harvesting in the first place. Two concepts are currently under debate: i) In 2013, Plenio et al. 31 argued that long-lived vibrationally coherent modes can significantly enhance the electronic coherence lifetime when the vibrational and elec- tronic degrees of freedom are resonantly coupled. The vibrational mode thereby is supposed to act as a "phonon laser" on the excitons, thereby producing ultralong electronic coherence. ii) More- over, Tiwari et al. 32 argued that nonadiabatic electronic-vibrational coherent mixing at short times resonantly enhance the amplitude of the particular delocalized anticorrelated vibrational mode on the ground electronic state. This second concept does not involve long-lived electronic coherence and is conceptually in agreement with the observation that a strong vibronic coupling produces large-amplitude coherent oscillations of the electronic component with a usual short lifetime and a long-lived vibrational coherence, but with a rather small amplitude 33, 34. This scenario was re- examined again recently 35 with an explicit coupling to an electronic and a vibrational bath. The result that an increased vibronic coupling survives weak electronic dephasing at short times and induces a resonantly enhanced long-lived vibrational coherence of the anticorrelated mode was presented as an explanation of the long-lived coherence signals observed in Ref. 6. However, 4 again, an unrealistically weak electronic damping with a reorganization energy of 35 cm−1 has been used. Motivated by this discrepancy, we have reexamined the coherent dynamics in a model of a vibronically coupled excitonic dimer with anticorrelated pigment vibrations. We calculate the 2D electronic spectra of the model dimer as being part of the FMO complex. We use the environmental parameters obtained from the recent FMO experiment 25. We examine the electronic coherence lifetime by the antidiagonal bandwith of the diagonal peak at zero waiting time. The oscillations in the residuals obtained from the global fitting analysis confirm that the long-lived coherence is purely vibrational in nature, irrespective of resonant or non-resonant conditions. To distinguish the coherent dynamics of the electronic excited state, we have calculated the dynamics of the electronic wave-packet on the vibrational potential energy surfaces (PESs), accompanied by the projection onto the reaction coordinates. Also here, the vibrational coherence is clearly identified by the oscillations close to the potential minimum. The projection shows that the long-lived oscillations are solely of vibrational origin, which confirms the 2D spectroscopic calculations. Moreover, we show that under realistically strong electronic damping, coherent vibronic coupling at short times does not enhance the amplitude of the anticorrelated vibrational mode, while we recover the mechanism of vibronic enhancement only for unrealistically weak electronic damping. 5 Theory The model is described by a total Hamiltonian consisting of the system, bath and the system-bath interaction terms, H = HS + HSB. The system is a dimer consisting of monomer A and B with site energies EA/B, both having the same electronic ground state g(cid:105) and the respective electronic single excited states A(cid:105) and B(cid:105). The double excited state is denoted as AB(cid:105). Each electronic excited state couples to a vibrational mode (each to its own mode). The two couplings are such that an anticorrelated out-of-phase oscillation of the two electronic states occurs. In the exciton site basis, we have HS = g(cid:105) hg (cid:104)g + A(cid:105) hA (cid:104)A + B(cid:105) hB (cid:104)B +(A(cid:105) V (cid:104)B + h.c.) + AB(cid:105) (hA + hB)(cid:104)AB . (1) Here, hg = 1 2Ω(P 2 A/B + Q2 A/B), hA = EA + hg − κQA and hB = EB + hg + κQB, respectively. PA/B and QA/B are the momenta and the coordinates of the two vibrational modes coupled to monomer A and B. We express the vibronic coupling strength between ground and excited state as κ = Ω∆√ 2 , where ∆ is the dimensionless shift of the excited state relative to its ground state. Ω is the vibrational frequency (both modes are taken with equal characteristics). V denotes the electronic coupling between two electronic excited states A(cid:105) and B(cid:105). For the discussion of the anticorrelated vibrations, it is useful to define new coordinates and (PA ± PB) 36. Then, the system (QA ± QB), and P± = 1√ momenta according to Q± = 1√ 2 2 6 Hamiltonian can be written as HS = (cid:88) HS,n = (cid:88) (H + S,n + H− S,n), n=0,1,2 n=0,1,2 H− S,0 = 10h−, √ 2)B(cid:105)(cid:104)B S,1 = 11h− + (EB − ΩQ−∆/ H− √ 2)A(cid:105)(cid:104)A +(EA + ΩQ−∆/ +V (A(cid:105)(cid:104)B + B(cid:105)(cid:104)A), H− S,2 = 12(h− + EA + EB) , √ S,n = 1n(h+ − nΩQ+∆/ (2) with [H + S,n] = 0 and H + 2Ω(P 2± + Q2±) and the projection operators are 1 =(cid:80) S,n, H− given by h± = 1 11 = A(cid:105)(cid:104)A + B(cid:105)(cid:104)B, and 12 = AB(cid:105)(cid:104)AB. 2). The two rotated vibrational modes are n=0,1,2 1n, with 10 = g(cid:105)(cid:104)g, We choose the same parameters as in Ref. 32 for the system Hamiltonian. This dimer mimics one exciton pair of the FMO complex. The bath part will be discussed below (in particular, we do not choose the same parameters of too weak damping, but use our own parameters of Ref. 25). The electronic energy gap is set to EA − EB = 150 cm−1 and the electronic coupling is V = 66 cm−1. Moreover, the dimensionless vibrational shift is set to ∆ = 0.2236. The pure electronic energy gap without coupling can then be calculated to be ∆E = 200 cm−1. As in Ref. 32, we model inhomogeneous broadening by a static Gaussian disorder of width δE = 26 cm−1. For the vibrational mode, we choose the frequency Ω = 200 cm−1 which corresponds to the resonant case when the vibronic coupling vanishes (the slight shift of this resonance due to the vibronic coupling does not alter the overall result since we find essentially the same conclusion also for the 7 off-resonant case, see below). The environment and coupling part HSB = H vib SB + H el SB consist of two parts, the vibrational baths which damp the vibrational motions, and the electronic bath which generates electronic de- phasing and damping. In general, we assume Gaussian fluctuations described in terms of the standard model of dissipative quantum systems. The electronic environment is generated by fluc- tuating charges in the protein and the solvent and consists of two harmonic oscillator baths each of which couples to the electronic excited states of monomer A and B, respectively. Thus, we have the Hamiltonian H el α=A,B H el SB,α with SB =(cid:80) (cid:34) N(cid:88) i=1 H el SB,α = 1 2 p2 i,α mi,α + mi,αω2 i,α (cid:18) xi,α − ci,α α(cid:105)(cid:104)α mi,αω2 i,α (cid:19)2(cid:35) . (3) As usual, pi,α and xi,α are the momenta and the coordinates of the ith bath mode coupling to the electronic state α = A, B. For the electronic part, we choose an Ohmic spectral density with the parameters obtained from fitting the linear spectra of the FMO complex to experimental data, see Ref. 25. Notice that these values correspond to much stronger damping than those in Ref. 32. Thus, each bath is assumed to have its own, but equal spectral density J el(ω) = γelω exp(−ω/ωc), with γel = 0.7, ωc = 350 cm−1. The vibrational environment roots in fluctuating nuclear degrees of freedom of the protein and couples to the vibrational displacements QA or QB of the mode coupled to the electronic state (cid:18) yi,α − di,αQα µi,αν2 i,α (cid:19)2(cid:35) . (4) A or B. Hence, H vib α=A,B H vib SB,α with SB =(cid:80) (cid:34) N(cid:88) i=1 H vib SB,α = 1 2 q2 i,α µi,α + µi,αν2 i,α 8 qi,α and yi,α are the momenta and the coordinate of the ith vibrational bath mode of the state α = A, B. We assume that the vibrational bath has the same spectral density as the electronic bath, i.e., J vib(ω) = γvibω exp(−ω/ωc) but with weaker damping, γvib = 0.02 and ωc = 350 cm−1. To disentangle electronic and vibrational coherence, we perform a projection of the electronic wave packet on the reaction coordinate, which allows us to distinguish the vibrational coherence from the vibronic dynamics. We assume the initial wave packet to be in the lowest vibrational state 0(cid:105) of the electronic excited state A(cid:105) in the site basis, such that the initial density matrix can be written as ρ(0) = A, 0(cid:105)(cid:104)A, 0. In order to obtain dynamical information, we determine the probability of the wave packet along the reaction coordinate Q− by the time-dependent projection k (Q−, t) = (cid:104)Q−(cid:104)k ρ(t)k(cid:105)Q−(cid:105) , P ad (5) where P ad k is the probability density of the reaction coordinate and k indicates the electronic state of A or B in the exciton basis (for details of the projection, see Refs. 37, 38). Results and Discussion We assume 32 that two perpendicular transitions from the common ground state to the two ex- cited states of monomer A and B are possible. Hence, the transition dipole moments are fixed to (cid:126)µA = µAex and (cid:126)µB = µBey with µA = µB = 1. Here, ej is the unit vectors in the direction j. Temperature is set to 300 K, if not stated otherwise. We use the time non-local quantum master equation 39 -- 41 with the equation-of-motion phase-matching-approach 42. Details are given in the 9 Supplementary Information Appendix of Ref. 25. We first consider the vibronic dimer with resonant vibrational coupling for which we obtain the 2D electronic spectrum shown in Fig. 1(A) (real part). At waiting time T = 0 fs, the inhomoge- neous broadening can be clearly identified because the spectrum is stretched along the diagonal. It disappears within 50 fs (see the time dependent 2D electronic spectra in the SI). The corresponding 2D spectrum of the purely electronic dimer without vibrational modes is shown in Fig. 1(B). From the profiles taken along the antidiagonal band and shown in Figs. 1 (C) and (D), we observe that the anti-diagonal broadening of the 2D spectra in both cases is similar, which proves that both dimers undergo a similar dephasing dynamics with similar time scales of the electronic dephasing. They are extracted to be 70 fs and 80 fs, respectively. To resolve the time-dependent energy transfer, we analyze the series of 2D spectra for increasing waiting times, by the global fitting approach, see Supplementary Information Appendix for details. This yields the shortest lifetime of the decay associated spectra which is induced by the peak broadening and by electronic dephasing. Both lifetimes coincide in both cases which shows that a vibronic coupling does not alter the short-time electronic dephasing properties. For a quantitative analysis of the dissipative dynamics in the presence of a vibrational cou- pling, we plot in Fig. 1 (E) the time evolution of the magnitude of the peaks selected in Fig. 1 (A). We observe that the dynamics can be clearly separated into two sectors: (i) fast electronic dephas- ing, which initially occurs on the time scale of ∼70 fs, as already resolved by the analysis of the anti-diagonal bandwidth and the global fitting approach. Moreover, (ii) long-lived oscillations with 10 small amplitudes are resolved at longer times. In order to identify the origin of these long-lived oscillations, we perform a Fourier transform of the residual, which is obtained by subtracting the kinetics resolved by the global fitting. The result is shown in Figs. 1 (F) and (G). The process of fast electronic dephasing is associated to the broad background spectral band with a maximum at 200 cm−1. In addition, this broad band is overlapped by one sharp peak at the same frequency of 200 cm−1. One additional narrow peak is resolved at 400 cm−1. It originates from the vibrational coherence between the vibrational ground 0(cid:105) and the second vibrational level 2(cid:105) of the electronic ground state, i.e., is of pure vibrational origin. This is further illustrated in Fig. S3 of the Sup- plementary Information Appendix, where the stick spectrum also indicates the clearly separated electronic and vibrational parts of which the eigenstates are composed. Hence, as a matter of fact, we can conclude that these narrow peaks at 200 cm−1 and 400 cm−1 only stem from vibrational coherence of each monomer. Off-resonance case Up to here, we have studied the vibronic dimer for the resonant case ∆E = Ω. Next, we investigate the off-resonant case as well and choose Ω = 500 cm−1. The results are shown in Fig. 2 (A), the global fitting analysis is shown in the Supplementary Information Appendix. The fast electronic dephasing with the time scale of 69 fs is still present. It agrees with the value of the antidiagonal bandwidth (see the SI). Importantly enough, it coincides with the dephasing time scale of the resonant case. Hence, the fact that the electronic and vibrational dynamics are off-resonant does not affect the conclusion reached for the resonant case. In addition, we show in Fig. 2 (B) the dynamics of the selected peaks for growing waiting times. It shows the same kinetics as in the resonant case: One fast electronic dephasing component is combined with a 11 long-lived vibrational coherent component with a small amplitude. Again, we perform the Fourier transform of the residuals and plot the spectra of each peak in Fig. 2 (C). We again find one broad band with a maximum at 200 cm−1, which manifests the fast electronic dephasing and coincides with the lifetime of ∼ 70 fs resolved by the global fitting approach. One clearly separated narrow peak is located at 500 cm−1 with a large magnitude which is associated to the long-lived vibrational coherence. A clear evidence for the purely vibrational (and not vibronic) origin of the peak is that one additional peak can be resolved at ∼ 1000 cm−1. It is the clear signature of the vibrational coherence between the vibrational ground 0(cid:105) and the second vibrationally-excited level 2(cid:105) on the electronic ground-state surface. Low-temperature case Next, we consider the case of low temperature of 80 K. We follow the same steps as above and find for vanishing vibronic coupling κ = 0 an electronic dephasing time scale of 161 fs, see Fig. S4 in the SI. In the presence of a resonant vibrational mode with Ω = 200 cm−1 and with κ (cid:54)= 0 set to the same value as above, we obtain the dephasing time of 121 fs, see Fig. S3 (B) in the SI. It is slightly smaller than the one of the purely electronic case, but still comparable. The dynamics of the selected peaks in the 2D spectra again shows long-lived oscillations, see Fig. S3 (C) of the SI. The Fourier transform, shown in Fig. S3 (D), shows again one sharp peak at 200 cm−1, and one additional peak at 400 cm−1 with a quite weak magnitude. They manifest again the vibrational origin of the coherence. Therefore, we can conclude that, also for low temperature, the long-lived oscillation is just of vibrational origin. No different mechanism between low and room temperature occurs. 12 Vibrational dynamics of the monomer In addition to the dimer, we also investigate the monomer where only vibrational coherence is present. In Fig. S5 (B), we show the time trace of the selected cross peak together with the Fourier spectrum in (C). The spectra are dominated by one peak at the vibrational frequency. An additional peak appears at the position of twice the vibrational frequency. Thus, the same scenario occurs for the monomer as well. We clearly demonstrate that the long-lived oscillations in a vibronically-coupled dimer are just due to the overlap of the short-lived electronic coherence and the long-lived vibrational coherence. Wave packet tracking A further confirmation of this picture is obtained from monitoring the dynamics of the electronic excited states. For this, we project the time-evolved density matrix onto the anticorrelated vibrational coordinate Q−. We use the same parameters as before and calculate the PESs of the electronic excited states A(cid:105) and B(cid:105) in the adiabatic basis. The result for the off-resonant case with Ω = 500 cm−1 is shown in Figs. 3 (A) to (D). The initial wave packet is prepared in the excited state A(cid:105). For growing time, the transfer of population from A(cid:105) to B(cid:105) can be clearly identified by the decrease of the magnitude in Fig. 3 (A) and the corresponding growth in Fig. 3 (B). By this, the vibrational coherence of the excited states is clearly visible from the oscillations around the potential minimum, see Fig. 3 (A). The oscillations have a period of ∼ 66 fs, which exactly coincides with the assigned vibrational frequency of 500 cm−1. Moreover, the population dynamics of the states A(cid:105) and B(cid:105) is shown in Fig. 3 (C) by summing the wave packet population along the reaction coordinate Q−. Spectral information can be again obtained from the Fourier transform. In Fig. 3 (D), the vibrational coherence is identified by the narrow peaks at 500 cm−1 and 1000 cm−1, which coincide with the results from the 2D spectroscopic calculations 13 shown in Fig. 2. In addition, a broadband background with a maximum at 200 cm−1 and with small magnitude is visible, which again provides evidence of the electronic coherence being short-lived. The resonant case, with Ω = 200 cm−1, is addressed in Fig. 3 (E) to (F). Compared to the off-resonant case, no significant difference occurs. The initial wave packet in the excited state A(cid:105) is transferred to B(cid:105) over time. The only difference is the vibrational oscillation period of ∼165 fs. The integrated time-dependent populations are shown in Fig. 3 (G) and the associated spectral information in Fig. 3 (H). One narrow peak at 200 cm−1 and one additional peak at 400 cm−1 with quite small magnitude occur. Also here, the result agrees with the observation of the 2D spectroscopic calculations in Fig. 1. Vibronic dimer under weak electronic dephasing Up to here, we have studied realistic parameters of the electronic dephasing and the vibrational damping constants. The possibility remains that for weaker electronic dephasing, the role of a coherent vibronic coupling could be more pronounced. That this is not the case follows from the dynamics of a vibronic dimer in the off-resonant case with Ω = 500 cm−1 under (unrealistically) weak electronic dephasing. For this, we set γel = γvib = 0.02, and ωc = 50 cm−1. The wave packet dynamics projected to the PESs of A(cid:105) and B(cid:105) is shown in Fig. 4 (A) and (B), respectively. The purely vibrational coherence can be seen from the wave packet oscillations around Q− = −1.5 with a period of ∼ 67 fs, which coincides with the vibrational period. The electronic coherence is visible in Fig. 4 as a large-amplitude population exchange between the two electronic states. The electronic oscillation period of ∼167 fs corresponds to the electronic energy gap ∆E = 200 cm−1 in the adiabatic basis. 14 Thus, the large-amplitude exchange is caused by the superposition of the wave packet components on the two PESs. To reveal the oscillation components and their lifetimes, we sum the wave packet components along the reaction coordinate and plot it in Fig. 4 (C). The Fourier spectrum is shown in Fig. 4 (D). Two large peaks at 200 cm−1 and 500 cm−1 correspond to the oscillations due to electronic and vibrational coherence, respectively. Two small vibronic peaks at 500 − 200 = 300 cm−1 and 500 + 200 = 700 cm−1 are due to the vibronic mixing. Most importantly, although the frequencies indeed mix and additional peaks are generated, the line widths of the peaks at 200 cm−1, 300 cm−1, 500 cm−1, and 500 cm−1 are 35 cm−1, 40 cm−1, 20 cm−1, and 45 cm−1, and are thus all comparable. This proves that the lifetime of the electronic coherence is not affected by vibronic coupling to a vibrational mode. Impact of coherent vibronic coupling on anticorrelated vibrations Finally, we address the possibility that a strong coherent vibronic coupling could enhance the amplitude of the anti- correlated component of the vibrational dynamics 32, 35. The latter is given by the magnitude of the vibrational peak in the Fourier spectrum of the wave-packet dynamics, as, e.g., shown in Figs. 3 D and H and Fig. 4D. In Fig. 5, we show the amplitude of the anticorrelated vibration for increasing vibronic coupling ∆ for weak (γel = 0.02) and strong (γel = 0.7) electronic dephasing. For in- creasing ∆, the mixing of the anticorrelated vibration with the electronic parts becomes stronger. Indeed, for weak electronic dephasing, we find an increase of the anticorrelated vibrational am- plitude which confirms the picture of Refs. 32, 35. However, for the more realistic case of stronger electronic dephasing, the amplitude of the anticorrelated vibration depends only weakly on the vi- bronic coupling, since the coherent electron-vibrational mixing is dephased very rapidly and does 15 not influence the vibration at later times. Conclusions In conclusion, we have shown that, under ambient physical conditions, it is irrelevant for the life- time of electronic quantum coherence in the excitation energy transfer whether the two exciton states couple to two anticorrelated or correlated vibrational modes. This holds irrespective of whether the electronic and the vibrational transitions are resonant or off-resonant and follows from an analysis of a model dimer in which each excited state is coupled to its own vibrational mode in an anticorrelated manner. Two independent baths for electronic dephasing as well as vibrational damping are included. By this, we answer a key question in the literature 31 whether a coupling to a long-lived vibrational mode can lead to a substantial increase of the electronic coherence time. The conclusions are drawn from the calculated dynamics, the 2D electronic spectra and the subsequent 2D global fitting approach. The exciton dynamics is characterized by the combination of a fast electronic dephasing and a long-lived vibrational coherent component, which has very small os- cillation amplitudes. The long-lived oscillations are solely due to the coherence between different vibrational levels, irrespective of a resonant or an off-resonant vibronic anticorrelated coupling. Even under (unrealistically) weak electronic dephasing, the electronic coherence lifetime is not enhanced by the vibronic components. The same conclusion has been drawn from the study of indocarbocyanine dye molecules 33. In addition, we find that a strong mixing of electronic and anticorrelated vibrational components of the wavefunction due to strong vibronic coupling does not enhance the vibrational amplitude at long times under ambient conditions. This effect only 16 can occur under unrealistically weak electronic dephasing 32, 35, but does not play a role in realistic physical systems, the reason being that the required coherent vibronic mixing is rapidly destroyed by fast electronic dephasing. 1. Blankenship RE (2014) Antenna complexes and energy transfer processes. Molecular Mech- anisms of Photosynthesis (Blackwell Science, Oxford/Malden, 2002), pp 61-94. 2. Jonas DM (2003) Two-dimensional femtosecond spectroscopy. Annu Rev Phys Chem 54:425- 463. 3. Cowan ML, Ogilvie JP, Miller RJD (2004) Two-dimensional spectroscopy using diffractive optics based phased-locked photon echoes. Chem Phys Lett 386:184-189. 4. Brixner T, Mancal T, Stiopkin IV, Fleming GR (2004) Phase-stabilized two-dimensional elec- tronic spectroscopy. J Chem Phys 121:4221-4236. 5. Mukamel S (1995) Nonlinear response functions and optical susceptibilities. Principles of Nonlinear Optical Spectroscopy (Oxford Univ Press, Oxford), pp 111-139. 6. Engel GS, et al. (2007) Evidence for wavelike energy transfer through quantum coherence in photosynthetic systems. Nature 446:782-786. 7. Panitchayangkoona G, et al. (2010) Long-lived quantum coherence in photosynthetic com- plexes at physiological temperature. Proc Natl Acad Soc USA 107:12766-12770. 8. Lambert N, Chen YN, Cheng YC, Li CM, Chen GY, Nori F (2013) Quantum biology. Nat Phys 9:10-18. 17 9. Collini E, Wong CY, Wilk KE, Curmi PMG, Brumer P, Scholes GD (2010) Coherently wired light-harvesting in photosynthetic marine algae at ambient temperature. Nature 463:644-647. 10. Schlau-Cohen GS, et al. (2012) Elucidation of the timescales and origins of quantum electronic coherence in LHCII. Nat. Chem. 4:389-395. 11. Fuller FD, Pan J, Gelzinis A, Butkus V, Senlik SS, Wilcox DE, Yocum CF, Valkunas L, Abra- mavicius D, Ogilvie JP (2014) Vibronic coherence in oxygenic photosynthesis. Nat Chem 6:706-711. 12. Romero E, Augulis R, Novoderezhkin VI, Ferretti M, Thieme J, Zigmantas D, van Grondelle R (2014) Quantum coherence in photosynthesis for efficient solar-energy conversion. Nat Phys 10:676-682. 13. Ishizaki A, Fleming GR (2009) Theoretical examination of quantum coherence in a photosyn- thetic system at physiological temperature. Proc Natl Acad Soc USA 106:17255-17260. 14. Plenio MB, Almeida J, Huelga SF (2013) Origin of long-lived oscillations in 2D-spectra of a quantum vibronic model: electronic versus vibrational coherence. J Chem Phys 139:235102. 15. Adolphs J, Renger T (2006) How proteins trigger excitation energy transfer in the FMO com- plex of green sulfur bacteria. Biophys J 91:2778-2797. 16. Wendling M, Pullerits T, Przyjalgowski MA, Vulto SIE, Aartsma TJ, van Grondelle R, van Amerongen H (2000) Electron-Vibrational Coupling in the Fenna-Matthews-Olson Complex of Prosthecochloris aestuarii Determined by Temperature-Dependent Absorption and Fluores- cence Line-Narrowing Measurements. J Phys Chem B 104:5825-5831. 18 Adolphs J, Renger T (2006) How proteins trigger excitation energy transfer in the FMO com- plex of green sulfur bacteria. Biophys J 91:2778-2797. 17. Chen L, Zheng R, Jing Y, Shi Q (2011) Simulation of the two-dimensional electronic spectra of the Fenna-Matthews-Olson complex using the hierarchical equations of motion method. J Chem Phys 134:194508. 18. Nalbach P, Braun D, Thorwart M (2011) Exciton transfer dynamics and quantumness of energy transfer in the Fenna-Matthews-Olson complex. Phys Rev E 84:041926. 19. Wu J, Liu F, Ma J, Silbey RJ, Cao JS (2012) Efficient energy transfer in light-harvesting systems: Quantum-classical comparison, flux network, and robustness analysis. J Chem Phys 137:174111. 20. Lee MK, Coker DF (2016) Modeling Electronic-Nuclear Interactions for Excitation Energy Transfer Processes in Light-Harvesting Complexes. J Phys Chem Lett 7:3171-3178. 21. Lee MK, Huo P, Coker DF (2016) Semiclassical Path Integral Dynamics: Photosynthetic En- ergy Transfer with Realistic Environment Interactions. Annu Rev Phys Chem 67: 639-668. 22. Oh SA, Coker DF, Hutchinson DAW (2018) Optimization of energy transport in the Fenna- Matthews-Olson complex via site-varying pigment-protein interactions. arXiv:1807.03459. 23. Chandrasekaran S, Aghtar M, Valleau S, Aspuru-Guzik A, Kleinekathofer U (2015) Influence of Force Fields and Quantum Chemistry Approach on Spectral Densities of BChl a in Solution and in FMO Proteins. J Phys Chem B 119:9995-10004. 19 24. Olbrich C, Strumpfer J, Schulten K, Kleinekathofer U (2011) Theory and Simulation of the Environmental Effects on FMO Electronic Transitions. J Phys Chem Lett 2:1771-1776. 25. Duan HG, et al. (2017) Nature does not rely on long-lived electronic quantum coherence for photosynthetic energy transfer. Proc Natl Acad Soc USA 114:8493-8498. 26. Egorova D (2014) Self-analysis of coherent oscillations in time-resolved optical signals. J Phys Chem A 118:10259-10267. 27. Butkus V, et al. Vibrational vs. electronic coherences in 2D spectrum of molecular systems. Chem Phys Lett 545:40-43. 28. Seibt J, Pullerits T (2013) Beating signals in 2D spectroscopy: electronic or nuclear coher- ences? application to a quantum dot model system. J Phys Chem C 117:18728-18737. 29. Kreisbeck C, Kramer T (2012) Long-lived electronic coherence in Dissipative exciton dynam- ics of light-harvesting complexes. J Phys Chem Lett 3:2828-2833. 30. Kreisbeck C, Kramer T, Aspuru-Guzik A (2013) Disentangling electronic and vibronic coher- ences in two-dimensional echo spectra. J Phys Chem B 117:9380-9385. 31. Chin AW, et al. (2013) The role of non-equilibrium vibrational structures in electronic coher- ence and recoherence in pigmentprotein complexes, Nat Phys 9:113-118. 32. Tiwari V, Peters WK, Jonas DM (2013) Electronic resonance with anticorrelated pigment vi- brations drives photosynthetic energy transfer outside the adiabatic framework. Proc Natl Acad Soc USA 110:1203-1208. 20 33. Duan HG, et al. (2015) On the origin of oscillations in two-dimensional spectra of excitonically-coupled molecular systems. New J Phys 17:072002. 34. Halpin A, et al. (2014) Two-dimensional spectroscopy of a molecular dimer unveils the effects of vibronic coupling on exciton coherences. Nat Chem 6:196-201. 35. Yeh S-H, Hoehn RD, Allodi MA, Engel GS, Kais S (2018) Elucidation of near-resonance vi- bronic coherence lifetimes by nonadiabatic electronic-vibrational state character mixing. Proc Natl Acad Soc USA, Article Ahead of Print, https://doi.org/10.1073/pnas.1701390115 36. Gelin MF, et al. (2012) Bath-induced correlations and relaxation of vibronic dimers. J Chem Phys 136:034507. 37. Manthe U, Koppel H (2009) Dynamics on potential energy surfaces with a conical intersection: Adiabatic, intermediate, and diabatic behavior. J Chem Phys 93:1658-1669. 38. Qi D, Duan HG, Sun ZL, Miller RJD, Thorwart M (2017) Tracking an electronic wave packet in the vicinity of a conical intersection. J Chem Phys 147:074101. 39. Meier C, Tannor DJ (1999) Non-Markovian evolution of the density operator in the presence of strong laser fields. J Chem Phys 111:3365-3376. 40. Kleinekathofer U (2004) Non-Markovian theories based on a decomposition of the spectral density. J Chem Phys 121:2505-2514. 41. Cheng YC, Fleming GR (2008) Coherence quantum beats in two-dimensional electronic spec- troscopy. J Phys Chem A 112:4254-4260. 21 42. Gelin MF, Egorova D, Domcke W (2005) Efficient method for the calculation of time- and frequency-resolved four-wave mixing signals and its application to photon-echo spectroscopy. J Chem Phys 123:164112. Acknowledgements We acknowledge financial support by the Max Planck Society and the Hamburg Centre for Ultrafast Imaging (CUI) within the German Excellence Initiative supported by the Deutsche Forschungsgemeinschaft. H-G.D. acknowledges generous financial support by the Joachim-Hertz-Stiftung Hamburg. Significance Statement We have studied the impact of molecular vibrations on the electronic coherence during the energy transfer in a model dimer. The full dynamics are revealed in the calculated 2D electronic spectra. We show that the long-lived coherence present in the off-diagonal spectral signals is solely due to vibrational coherence in the monomer. Our calculations illustrate that neither the electronic coherence between two monomers can be enhanced by vibrations of individual pigments, irrespective of resonant or off-resonant conditions, nor can a coherent vibronic coupling enhance the amplitude of the anticorrelated vibrational mode under realistic conditions. Supporting Information The Supplementary Information includes the global fitting approach and results, low-temperature calculations and vibrational dynamics of monomer. Competing Interests The authors declare that they have no competing financial interests. Correspondence [email protected] and [email protected] 22 Figure 1: (A) Real part of the 2D electronic spectrum of the vibronic dimer at room temperature (300 K). The diagonal peak A is located at (ωt = 150 cm−1, ωτ = 150 cm−1) and B at (ωt = −50 cm−1, ωτ = −50 cm−1). The off-diagonal peak C sits at (ωt = −50 cm−1, ωτ = 150 cm−1) and D at (ωt = 150 cm−1, ωτ = −50 cm−1). The time dependent trace of the selected peaks are shown in (E). For comparison, the 2D spectrum of the dimer without the effective vibrational mode is shown in (B). To obtain the timescale of the electronic dephasing, the anti-diagonal profile of the peaks B in (A) and (B) are shown in (C) and (D). The resulting timescale of the electronic dephasing is 70 fs and 80 fs, respectively. (F), (G) Power spectra of the peaks A, B, C, and D for the case with vibrational coupling. The broad background spectral band is associated to fast electronic dephasing. In addition to the strong vibrational peak at 200 cm−1, one additional peak at 400 cm−1 is well resolved. 23 Figure 2: (A) Real part of the 2D electronic spectrum of the vibronic dimer under off-resonant conditions with Ω = 500 cm−1 at T = 0 fs. The kinetics of the selected peaks at A, B, C, and D are shown in (B), the associated power spectra are shown in (C). To resolve the lifetime of the electronic coherence, we fit the broad peak at 200 cm−1 with the Lorentzian lineshape and obtain the electronic coherence time of ∼70 fs, which agrees with the resonant case. 24 Figure 3: Time evolution of the wave-packet in the excited states A(cid:105) (A) and B(cid:105) (B). The integrated populations of the electronic states A(cid:105) and B(cid:105) obtained by summing along the reaction coordinate Q− are shown in (C), together with the Fourier transform of the residuals in (D). The long-lived vibrational coherence is identified by the narrow peaks at 500 cm−1 and 1000 cm−1. In addition, one broadband peak at 200 cm−1 represents short-lived electronic coherence. The corresponding results of the resonant case are shown in (E) to (H), respectively. The long-lived vibrational coherence can be identified by the narrow peak and 200 cm−1 and the small peak at 400 cm−1. 25 Figure 4: Time evolution of the wave-packet on the excited state PESs for A(cid:105) is shown in (A) and for B(cid:105) in panel (B) for the vibronic dimer under off-resonant conditions for very weak electronic dephasing γel = γvib = 0.02, and ωc = 50 cm−1. The integrated populations of the electronic states A(cid:105) and B(cid:105) obtained by summing along the reaction coordinate Q− are shown in (C), together with the Fourier transform of the residuals shown in (D). The vibronic coherence can be identified by the two peaks 300 cm−1 and 700 cm−1, which are marked by circles. 26 Figure 5: Oscillation amplitude of the anticorrelated vibration vs. the vibronic coupling strength ∆ for weak (γel = 0.02, ωc = 50 cm−1) and strong (γel = 0.7, ωc = 350 cm−1) electronic dephasing for ∆E = 200 cm−1, Ω = 500 cm−1, and T = 300 K. 27 00.10.20.30.40.50.60123456x 104∆Amplitude (arb. units) weak dampingstrong dampingfit 1fit 2
1512.07611
1
1512
2015-12-23T20:10:10
Mushroom spore dispersal by convectively-driven winds
[ "physics.bio-ph" ]
Thousands of fungal species rely on mushroom spores to spread across landscapes. It has long been thought that spores depend on favorable airflows for dispersal -- that active control of spore dispersal by the parent fungus is limited to an impulse delivered to the spores to carry them clear of the gill surface. Here we show that evaporative cooling of the air surrounding the mushroom pileus creates convective airflows capable of carrying spores at speeds of centimeters per second. Convective cells can transport spores from gaps that may be only a centimeter high, and lift spores ten centimeters or more into the air. The work reveals how mushrooms tolerate and even benefit from crowding, and provides a new explanation for their high water needs.
physics.bio-ph
physics
Mushroom spore dispersal by convectively-driven winds E. Dressaire,1 Lisa Yamada,2 Boya Song,3 and Marcus Roper3, 4 1New York University Polytechnic School of Engineering, Brooklyn, NY 11201 2Dept. of Engineering, Trinity College, CT 06106 3Dept. of Mathematics, UCLA, Los Angeles, CA 90095, USA 4Dept. of Biomathematics, UCLA, Los Angeles, CA 90095, USA Thousands of fungal species rely on mushroom spores to spread across landscapes. It has long been thought that spores depend on favorable airflows for dispersal – that active control of spore dispersal by the parent fungus is limited to an impulse delivered to the spores to carry them clear of the gill surface. Here we show that evaporative cooling of the air surrounding the mushroom pileus creates convective airflows capable of carrying spores at speeds of centimeters per second. Convective cells can transport spores from gaps that may be only a centimeter high, and lift spores ten centimeters or more into the air. The work reveals how mushrooms tolerate and even benefit from crowding, and provides a new explanation for their high water needs. layer of nearly still air, with typical thickness δ ∼(cid:113) νL U [15], where U is the horizontal wind velocity, L the size of the pileus, and ν the viscosity of air [15]; no external airflow can penetrate into gaps narrower than 2δ. For typically sized mushrooms under a grass canopy (with L = 10 cm, U = 1 − 10 cm/s, ν = 1.5 × 10−5 m2s−1), we find that 2δ ≈ 6 − 20 mm. If the gap thickness between pileus and ground is smaller than 2δ then no external wind will penetrate into the gap. RESULTS AND DISCUSSION Spores can disperse from thin gaps beneath pilei without external winds. We analyzed spore de- position beneath cultured mushroom (shiitake; Lentin- ula edodes, and oyster; Pleurotus ostreatus, sourced from CCD Mushroom, Fallbrook, CA) as well as wild-collected Agaricus californicus. Pilei were placed on supports to create controllable gap heights beneath the mushroom, and placed within boxes to isolate them from external airflows. We measured spore dispersal patterns by allow- ing spores to fall onto sheets of transparency film, and photographing the spore deposit (Fig. 1C). In all ex- periments, spore deposits extended far beyond the gap beneath the pileus. Spores were deposited in asymmet- ric patterns, both for cultured and wild-collected mush- rooms (Fig. 1D). Using a laser light sheet and high speed camera (see Materials and Methods), we directly visual- ized the flow of spores leaving the narrow gap beneath a single mushroom (Fig. 1E and Movie S1): our videos show that spores continuously flow out from thin gaps, even in the absence of external winds. What drives the flow of spores from beneath the pileus? Some ascomycete fungi and ferns create dispersive winds by direct transfer of momentum from the fruiting body to the surrounding air. For example, some ascomycete fungi release all of their spores in a single puff; the mo- mentum of the spores passing through the air sets the 5 1 0 2 c e D 3 2 ] h p - o i b . s c i s y h p [ 1 v 1 1 6 7 0 . 2 1 5 1 : v i X r a INTRODUCTION Rooted in a host organism or patch of habitat such as a dead log, tens of thousands of species of filamentous fungi rely on spores shed from mushrooms and passively carried by the wind to disperse to new hosts or habitat patches. A single mushroom is capable of releasing over a billion spores per day [1], but it is thought that the prob- ability of any single spore establishing a new individual is very small [2, 3]. Nevertheless in the sister phylum of the mushroom-forming fungi, the Ascomycota, fungi face similarly low likelihoods of dispersing successfully, but spore ejection apparatuses are highly optimized to maximize spore range [4–6], suggestive of strong selec- tion for adaptations that increase the potential for spore dispersal. Spores disperse from mushrooms in two phases [7]: a powered phase, in which an initial impulse delivered to the spore by a surface tension catapult carries it clear of the gill or pore surface, followed by a passive phase in which the spore drops below the pileus and is carried away by whatever winds are present in the surrounding environment. The powered phase requires feats of engi- neering both in the mechanism of ejection [8–10] and in the spacing and orientation of the gills or pores [11, 12]. However, spore size is the only attribute whose influ- ence on the passive phase of dispersal has been studied [13]. Spores are typically less than 10µm in size, so can be borne aloft by an upward wind of only 1 cm/s [11]. Buller claimed that such wind speeds are usually attained beneath fruiting bodies in Nature [11]: Indeed peak up- ward wind velocities under grass canopies are of order 0.1-1 cm/s [14]. However even if the peak wind veloc- ity in the mushroom environment is large enough to lift spores aloft: 1. the average vertical wind velocity is zero, with intervals of downward as well as upward flow and 2. mushrooms frequently grow in obstructed environments, such as close to the ground or with pilei crowded close together (Fig 1 A-B). The pileus traps a thin boundary air into motion [4]. Fern sporangia form over-pressured capsules that rupture to create jets of air [16]. However, the flux of spores from a basidiomycete pileus is thou- sands of times smaller than for synchronized ejection by a ascomycete fungus, and pilei have no known mechanism for storing or releasing pressurized air. The only mech- anism that we are aware of for creating airflows without momentum transfer is by the manipulation of buoyancy – an effect that underpins many geophysical flows [17] and has recently been tapped to create novel locomotory strategies [18]. Mushrooms evaporatively cool the surrounding air. Many mushrooms are both cold and wet to the touch [19]; the expanding soft tissues of the fruit body are hydraulically inflated, but also lose water quickly (Fig. 2A). We made a comparative measurement of water loss rates from living mushrooms and plants. The rate of water loss from mushrooms greatly exceeds water loss rates for plants, which use stomata and cuticles to limit evaporation (Fig. 2B). For both plants and pilei, rates of evaporation were larger when tissues were able to ac- tively take in water via their root-system / mycelium, than for cut leaves or pilei. However, the pilei lost water more quickly than all species of plants surveyed under both experimental conditions. In fact, evaporation rates from cut mushrooms were comparable to a sample of wa- ter agar hydrogel (1.5% wt/vol agar), while evaporation rates from mushrooms with intact mycelia, were twofold larger (Fig 2B). Taken together, these data suggest that pilei are not adapted to conserve water as effectively as the plant species analyzed. The high rates of evaporation lead to cooling of the air near the mushroom, and may have adaptive advantage to the fungus. Previous observations have shown that evap- oration cools both the pileus itself and the surrounding air by several degrees Celsius [20]. Specifically, since la- tent heat is required for the change of phase from liquid to vapor, heat must continually be transferred to mush- room from the surrounding air. We compared the ambi- ent temperature of the air between 20 cm and 1 m away from P. ostreatus pilei with intact mycelia, with tempera- tures in the narrow gaps between and beneath pilei, using a Traceable liquid/gas probe (Control Company, Forest- wood, TX). We found that gap temperatures were consis- tently 1-2◦C cooler than ambient (Fig. 2C). The surface temperature of the pileus, measured with a Dermatemp infra-red thermometer (Exergen, Watertown, MA) was up to 4◦C cooler than ambient, consistent with previous observations ([20] and Fig. 2C). Evaporation alone can account for these temperature differences: in a typical ex- perimental run, a pileus loses water at a rate of 3× 10−5 kg/m2s (comparable with the data of [21]). Evaporating this quantity of water requires Evap ≈ 70 W/m2 of va- porization enthalpy. At steady state the heat flux to the mushroom must equal the enthalpy of vaporization; New- ton’s law of cooling gives that the heat flux (energy/area) 2 will be proportional to the temperature difference, ∆ be- tween the surface of the mushroom and the ambient air: Evap = h∆T where h = 10 − 30 W/m2 oC is a heat transfer coefficient [22]. From this formula we predict that ∆T ≈ 2.5 − 7◦C, in line with our observations. ρ0 von Karman’s law: Ug ∼(cid:113) 2∆ρT gh Increasing air density by cooling produces dis- persive currents. Cooling air from the laboratory am- bient (T = 18◦C) down to T = 16◦C increases the density of the air by ∆ρT = α∆T = 0.008 kg m−3, where α is the coefficient of expansion of air. Cold dense air will tend to spread as a gravity current [23], and an order of magnitude estimate for the spreading velocity of this gravity current from a gap of height h = 1 cm is given by ≈ 4cm/s. Although the air beneath the pileus is laden with spores, spore weight contributes negligibly to the creation of disper- sive winds: in typical experiments, spores were released from the pileus at a rate of q =540± 490 spores/cm2 s (data from 24 P. ostreatus mushrooms). If the mass of a single spore is ms = 5 × 10−13kg and its sedimentation speed is vs = 10−3 m/s (see discussion preceding Equa- tion (1)), then the contribution of spores to the density of air beneath the pileus is ∆ρS = msq/vs = 0.003 kg m−3, less than half of the density increase produced by cool- ing ∆ρT . Indeed, water evaporation, rather than spores and the water droplets that propel them [11], constitute most of the mass lost by a mushroom. To prevent spore ejection, we applied a thin layer of petroleum jelly to the gill surfaces of cut mushrooms. Treated mushrooms lost mass at a statistically indistinguishable rate to cut mushrooms that were allowed to shed spores (Fig. 2B). Although our experiments were performed in closed containers to exclude external airflows it is still possi- ble that spore deposit patterns were the result of convec- tive currents created by temperature gradients in the lab, rather than airflows created by the mushroom itself. To confirm that spores were truly dispersed by airflows cre- ated by the mushroom we rotated the mushroom either 90◦ or 180◦ halfway through the experiment and replaced the transparency sheet. Since the box remained in the same orientation and position in the lab, we would ex- pect that if spores are dispersed by external airflows, the dispersal pattern would remain the same relative to the lab. In fact we found consistently that the direction of the dispersal current rotated along with the mushroom (see Figure S1), indicating that mushroom generated air- flows are dispersing spores. We explored how the distance dispersed by spores de- pended on factors under the control of the parent fungus. The distance spores dispersed from the pileus increased in proportion to the square of the thickness of the gap be- neath the pileus (R2 = 0.90, Fig. 3). However, we found no correlation between spore dispersal distance and the diameter of the pileus or the rate at which spores were produced (R2 = 0.07 and R2 = 0.17 respectively, Figure S2). Spores were typically deposited around mushrooms in asymmetric patterns, suggesting that one or two tongues of spore laden air emerge from under the pileus, and spores do not disperse symmetrically in all directions (Fig. 1C). These tongues of deposition were seen in wild- collected as well as cultured mushrooms (Fig. 1D). We dissected the dynamics of one of these tongues by build- ing a two dimensional simulation of the coupled temper- ature and flow fields around the pileus. Although real dispersal patterns are three dimensional, these simula- tions approximate the 2D dynamics along the symmetry plane of a spreading tongue. In our simulations we used a Boussinesq approximation for the equations of fluid motion and model spores as passive tracers, since their mass contributes negligibly to the density of the grav- ity current. Initially we modeled the pileus by a perfect half-ellipse whose diameter (4 cm) and height (0.8 cm) matched the dimensions of a L. edodes pileus used in our experiments. However, if cooling was applied uniformly over the pileus surface then spores dispersed weakly (Fig. 4A). Weak symmetric dispersal can be explained by con- servation of mass: cold outward flow of spore-laden air must be continually replenished with fresh air drawn in from outside of the gap. In a symmetric pileus, the cool air spreads along the ground and inflowing air travels along the under-surface of the pileus. So initially on leav- ing the gills of the mushroom, spores are drawn inward with the layer of inflowing warm air; and only after spores have sedimented through this layer into the cold outflow beneath it do they start to travel outward (Fig 4A, upper panel). Asymmetric airflows are necessary for disper- sal. To understand how mushrooms can overcome the constraints associated with needing to maintain both in- flow and outflow, we performed a scaling analysis of our experimental and numerical data. Although the buoy- ancy force associated with the weight of the cooled air draws air downward, warm air must be pulled into the gap by viscous stresses. For fluid entering a gap of thick- ness h, at speed U , the gradient of viscous stress can be estimated as: ∼ ηU/h2, where η is the viscosity of air. We estimate the velocity U by balancing the vis- cous stress gradient with the buoyancy force; ρgα∆T i.e.: U ∼ ρgα∆T h2/η. We then adopt the notation that if f is a quantity of interest (e.g. temperature or gap height) that can vary over the pileus, then we write fr for the value of f on the right edge of the pileus and f(cid:96) for its value on the left edge of the pileus. If Ur = U(cid:96) then there is the same inflow on the left and right edges of the pileus, and dispersal is symmetric and weak. If there are different inflows on the left or right side of the pileus, then there can be net unidirectional flow beneath the pileus, carrying spores further. Assuming, without loss of generality, that the net dispersal of spores is rightward, the spreading velocity of the gravity cur- 3 rent can be estimated from the difference: Ug ∼ U(cid:96) − Ur (right-moving inflow minus left-moving inflow). The fur- thest traveling spores originate near the rightward edge of the pileus and fall a distance hr (the gap width on the rightward edge) before reaching the ground. Since the gravity current spreads predominantly horizontally, the vertical trajectories of spores are the same as in still air, namely they sediment with velocity vs and take a time ∼ hr/vs to be deposited. By balancing the weight of a spore against its Stokes drag, we obtain vs = 2ρa2g where ρ = 1.2 × 103 kg/m3 is the density of the spore, and a = 2 − 4 µm is the radius of a sphere of equivalent volume[4]. The sedimentation velocity, vs, can vary be- tween species, (typically vs = 1 − 4 mm/s), but does not depend on the flow created by the pileus. The maximum spore dispersal distance is then: 9η (cid:2)∆T h2(cid:3)(cid:96) r dmax = Ughr vs ∼ gα η hr vs (1) rh3 rh2 r, [h]l r (and Ug ∝ ∆T [h](cid:96) where we use the notation [f ]l r to denote the difference in the quantity f between the left and right sides of our model mushroom. Unidirectional dispersal therefore re- quires either that ∆T(cid:96) (cid:54)= ∆Tr (i.e. there is a tempera- ture gradient between the two sides of the pileus, Fig 4A middle panel) or h(cid:96) (cid:54)= hr (i.e. the mushroom is asym- metrically shaped, Fig 4A lower panel). These two cases can be distinguished by the dependence of dispersal dis- tance upon the gap height: for a temperature-gradient induced asymmetry we predict: dmax ∝ [∆T ](cid:96) r (and Ug ∝ [∆T ](cid:96) rh2 r), whereas for a shape induced asymmetry: dmax ∝ ∆T [h](cid:96) rhr). Both scalings are validated by numerical simulations (Fig. 4B). When rescaled using Equation (1), data from simulations with different values of h(cid:96), hr, [∆T ]l r or vs all collapsed to two universal lines corresponding to temperature and height asymmetries (Fig. 4C,D). Experimental data from real mushrooms is fit well by dmax ∼ h2 r (Fig. 3). Additionally the scaling (1) accords with experimental observations that dispersal distance does not depend on pileus diameter or on the rate of spore release (Fig. S2). We can not directly confirm that there are not temperature gradients over the pileus surface. But these data are consistent with shape asym- metry; that is variation in the thickness of the pileus, and height of the gap, playing a dominant role in asymmet- ric spore dispersal from real mushrooms. Additionally, we made another direct test of our scaling law (1) by measuring Ug (Fig. 4E, and see Materials and Methods). When the dimensionless prefactors in (1) were kept con- stant by using the same pileus, but the height of the gap, hr, was varied, we found that Ug ∼ hr, consistent with the scaling for shape asymmetry-driven flow (Fig. 4F). Spores can disperse over barriers surrounding the pileus. In nature, pilei may grow crowded together or under plant litter or close to the host or substrate con- taining the parental mycelium (Fig 1A-B); thus, in addi- tion to needing to disperse from the narrow gap beneath the pileus, spores may potentially also need to climb over barriers to reach external airflows. Although the cold, spore laden air is denser than the surrounding air, to re- place the cold air that continuously flows from beneath the pileus, warm air must be drawn in to the pileus. Our simulations showed that when the gravity current met a solid barrier, the warm inflow and cold outflow could link to form a convective eddy (Figure S3). We stud- ied whether this eddy could lift spores into the air, and what effect this might have upon spore dispersal. We found that when mushrooms were surrounded by a ver- tical barrier (see Materials and Methods), spores were dispersed over the barrier (Fig. 5A) provided that their horizontal range, predicted using Equation (1), exceeded the height of the barrier (Fig. 5B). Surprisingly, spores climbing over the barrier were dis- persed apparently symmetrically in all directions from the top of the barrier (compare Fig. 5A with 1C) over the entire area of the box containing the mushroom and bar- rier (Fig. 5A). In particular, the total horizontal extent of the spore deposit (the size of the box) typically greatly exceeded the horizontal range predicted by Equation (1), even without accounting for the distance traveled by the gravity current in crossing over the barrier. The enhancement of spore dispersal by a barrier can be attributed to the action of the recirculating eddy: spores that climb the barrier may enter the current of air that is pulled down to the mushroom to replace the cold spore laden air (Fig. 5C). This eddy can carry spores up and further away from the barrier, and it is likely that no longer being constrained to travel along the ground sur- rounding the mushroom contributes to their increased range. Why does the height that gravity currents need to climb up the barrier not reduce their dispersal distance? In a climbing gravity current spores continue to sedi- ment vertically downward. Since the vertical velocity of a climbing gravity current (Ug ∼ 4 cm/s) is typi- cally much larger than the sedimentation speed (vs ∼ 1 mm/s), spores do not sediment out of vertical gravity current as it climbs. To test this idea quantitatively, we numerically simulated spore dispersal by a gravity cur- rent that encounters a barrier inclined at an angle θ to the horizontal. Since the velocity of the gravity current is directed parallel to the barrier, spores can still sedi- ment toward the barrier, but at a reduced sedimentation velocity vs cos θ. (Sedimentation parallel to the barrier, with velocity vs sin θ can be neglected). Revisiting Equa- tion (1) we therefore predict that spores will disperse a distance Ughr/(vs cos θ), i.e. they will travel a factor of 1/ cos θ further up the slope than they would travel hor- izontally, and this prediction agrees quantitatively with numerical simulations (Fig. 5C). In particular, if θ = 90o (a vertical wall), we predict no sedimentation onto the 4 wall, consistent with experimental observations that the distance that a gravity current climbs up a vertical wall does not reduce its horizontal spreading. Finally we note that, though our experiments were de- signed to exclude external winds, in nature wind speed tends to increase with height above the ground [14], so spore that travel upward and then away from the vertical wall may be more likely to reach dispersive winds. CONCLUSION Mushrooms are not simple machines for producing the largest number of spores, but directly influence the dispersal of those spores, even during the phase of their dispersal previously considered to be passive. Rather than being the result of failure to evolve water- conserving adaptations, rapid water loss from the pileus enables mushrooms to create convective cells for dis- persing spores. Convectively-generated winds provide a mechanism for spore dispersal for pilei that are crowded close together or close to the ground. Indeed, the pres- ence of nearby boundaries for the upward flowing part of the current to climb may enhance spore dispersal. Our analysis of convective dispersal adds to an increas- ing body of work revealing that fungal sporocarps are exquisitely engineered to maximize spore dispersal poten- tial [5, 6, 24]. Although evaporative cooling is an essential ingredient for spore dispersal and has been observed in many species [20], our analysis reveals the previously un- reported role played by the asymmetric thickness of the pileus, or of the gap beneath it in shaping and amplifying the dispersive wind created by the mushroom. Asym- metric spreading of spores was seen both in cultured and wild-collected mushrooms (respectively L. edodes and A. californicus) with circular pilei. Although in some environments mushrooms may be able to rely on external winds to disperse their spores from the moment that spores are ejected, we have high- lighted two common growing configurations in which there is likely to be little external wind in the gap be- neath the pileus. There are an unknown number of such low-wind environments or niches for which the pileus gen- erated wind may drive or assist spore dispersal. More- over, the combination of controls that can be used by the fungus to manipulate dispersal distance (including gap width, asymmetry and temperature gradients) cre- ates the potential for mechanistic explanations for why some species may disperse more effectively than others [13, 25]. MATERIALS AND METHODS Evaporation rate We compared rates of water loss between L. edodes and P. ostreatus mushrooms and plants (one potted specimen of each of: Tagetes erecta, Ocimum basilicum, Salvia elegans, Chrysanthemum sp, Dahlia variabilis, acquired from Home Depot, cut leaves of Cerces canadensis, Pittosporum tobira, Tagetes erecta and Trachelospermum jasminoides, collected on the UCLA campus). To measure the rate of evaporation from mushrooms with intact mycelia or from whole plants we wrapped the pot or the log on which the mushroom grew in plastic, so that water was lost only through the pileus / leaves. We also compared with water agar (1.5% wt/vol) samples that were poured into petri dishes, and left un- covered. All samples were allowed to dry in a laboratory ambient (19◦C, 60% R.H.). We measured surface area of samples by cutting the leaves and pilei into pieces and photographing the pieces. Dispersal distance. For spore dispersal measurements mushroom caps were placed in closed containers to iso- late them from surrounding airflows. Depending on the gap height and length of gravity current, we used either 15cm diameter suspension dishes, food storage contain- ers to which we added cardboard drop-ceilings, or file boxes. To prevent condensation from forming under the mushroom pileus we added a small quantity (∼0.2 g) of desiccant (DampCheck, Orlando, FL) to each container. Mushroom gap height was controlled by balancing the mushrooms on skewers spanning between two cardboard trestles, short lengths of tubing or wire connectors, or by suspending them from the ceiling of the container. We measured both the distance dispersed by the spores, and the asymmetry of the spore print. To calculate the print asymmetry in Figure 1D, we traced, using ImageJ, the coordinates of the boundary of the spore print. We then calculated asymmetry moments: cn = (cid:82) eniφ dA, where . c0 c1+c2 the integral is carried out over the area of the spore print, and φ is the angle made between the line joining each point to the centroid of the originating pileus and an arbitrary orientation axis. We characterized the asym- metry using a parameter Asym = Velocimetry. Spore dispersal was directly observed by illuminating the pileus with a laser sheet created by ex- panding a vertical laser beam from a 0.5W Hercules- 450 laser (LaserGlow, Toronto, Canada) with a plano- concave lens (f=-7.7 mm, ThorLabs, Newton, N.J.). The mushroom pileus was placed inside a 18×22×22mm box: the ceiling and three walls of the box were lined with matte black aluminum foil (ThorLabs), and the fourth wall constructed of transparent acrylic, the floor of the container was covered with photography velvet to mini- mize scatter from the laser. The laser light sheet passed through a slit in the ceiling of the box to illuminate the rim of the mushroom cap. Spores traveling through the plane of the light sheet acted as bright tracer particles of the flow, and we filmed their movement using a FAST- CAM SA3 high speed camera (Photron, San Diego, CA). Because of laser heating, after 2-4 s the laser created a thermal plume that obscured the airflows created by the 5 mushroom. Accordingly, we analyzed only the first 2 s of video captured from each experiment, and waited 5 min to allow the surface to return to temperature equilibrium between runs of the experiment. Although our observa- tion time was necesarily short, the time between obser- vations was long enough for even slow overturning eddies to emerge. Spore velocities were measured by using a hy- brid of particle imaging velocimetry (based on the code of [26]) to measure the velocities of groups of spores and individual particle tracking [27]. Barrier crossing experiments To measure the ability of mushroom-created airflows to disperse spores over bar- riers, we surrounded mushrooms by cylindrical walls, formed by taping a strip of transparency sheet inside of a 15cm diameter petri dish (Fig. 5A). By varying the width of the strip, we could vary the height of the bar- rier. To measure whether spores could cross this barrier, we placed transparency sheet on the floor of the petri dish (inside the barrier) and on the bench top outside of the barrier. To measure spore deposition just inside, and just outside of the barrier, we cut annuli of diameter 15mm from the two transparency sheets. The two annuli were cut into smaller pieces, and then separately vortexed in 10ml deionized water to wash off spores. The concen- tration of spores from inside and outside the barrier was counted automatically by the following method: Spore suspensions were added to a hemocytometer to create samples with known volume, which were photographed at 260× magnification using a Zeiss AxioZoom micro- scope. Spores appear as dark ellipses with bright out- lines. Images were first denoised using median filtering and contrast enhanced using a Laplacian of Gaussian fil- ter. Images were thresholded to identify bright edges and dark spore interior regions. Spores were segmented mor- phologically, by first removing all edges shorter than 30 pixels, and all dark regions with diameter smaller than 20 pixels or larger than 80 pixels. Spores were identi- fied by dilating the dark regions using a disk element with a radius of 5 pixels, and keeping only those regions that overlapped with a detected bright edge. We then subtracted the bright edges off from the dark regions, to create one disconnected region per spore. The number of disconnected regions in each image was then counted. The ratio of the number of spores in the outer annulus to the number in the inner annulus gives a measure of the fraction of spores dispersed over the barrier. Since there is much larger total area of spore fall outside of the bar- rier than inside, the barrier crossing rate is proportional to, but not equal to, the fraction of spores that crossed the barrier. Numerical simulations To simulate the trajectories of spores in convective currents, we used the Boussinesq ap- proximation. Specifically to calculate the inertia of the gravity current we take the density of air to be constant ρ0, but when calculating the buoyancy force, we use a lin- ear model: ρ(T ) = ρ0−ρ0αT , where T is the temperature measured relative to the ambient temperature far from the mushroom (so that T → 0 far from the pileus), and α is a coefficient of expansion: α = 3.42 × 10−3/K. We non-dimensionalize our equations by scaling all distances by L, the half-width of the mushroom, all temperature differences by ∆T , the maximum temperature drop at the mushroom surface, velocities by U∗ = (αgL∆T )1/2 and pressure by ρ0U∗2. The temperature and velocity fields around the pileus satisfy coupled partial differen- tial equations (PDEs): ∇2u + T 1 Re u · ∇u = −∇p + ∇ · u = 0 u · ∇T = 1 P ´e ∇2T (2) Spore trajectories were calculated by solving the ODEs: x = u, y = v−vs/U∗, where vs is the spore sedimentation speed. Equations (2) contain two dimensionless numbers; the Reynolds number Re ≡ ρ0U∗L/η, formed from our velocity and length scales and the viscosity η of air and the P´eclet number P ´e ≡ U∗L/κ which depends on the thermal diffusivity, κ. In a typical simulation Re = 60 and P ´e = 40. We used Comsol Multiphysics (COMSOL, Los Angeles) to set up and solve the PDEs in a 2D do- main, in which the pileus was modeled by a semi-ellipse with a no-slip, constant temperature boundary condition on its surface. The pileus was set a small distance above a no-slip floor, and the domain was closed by a semi-circle on which the no-slip boundary condition was imposed, and whose diameter is 40× larger than the cap diam- eter. Constant temperature boundary conditions were applied on all surfaces, with T = 0 on the walls of the do- main, and cooling applied on the pileus surface (see main text for the two principal boundary conditions used). To avoid creating convective over-turning of fluid in the gap beneath the pileus, we isolated a section of the floor di- rectly beneath the mushroom, and applied a zero-flux boundary condition there. To model the effect of nearby walls, we replaced the semi-circular external boundary by boundaries that were set a distance 1.5 from the mid- line of the mushroom, and oriented at an angle θ to the horizontal, as described in the main-text. We thank Mechel Henry, Clive Roper, Christine Roper and Junius Santoso for experimental assistance and Mike Lawrence from UCLA Laser Safety for assistance with ex- perimental design. L.Y. was supported by the Southern California Applied Math REU program (DMS-1045536). MR is supported by a fellowship from the Alfred P. Sloan Foundation, by NSF grant DMS-1312543 and by set-up funds from UCLA. E.D. is supported by set-up funds by NYU Polytechnic School of Engineering. We thank CCD Mushroom Inc. for providing the mushroom logs used in this experiment. 6 [1] Kadowaki K, Leschen RA, Beggs JR (2010) Periodicity of spore release from individual ganoderma fruiting bodies in a natural forest. Australasian Mycologist 29:17–23. [2] Nagarajan S, Singh DV (1990) Long-distance disper- sion of rust pathogens. Annual review of phytopathology 28:139–153. [3] Galante TE, Horton TR, Swaney DP (2011) 95% of basidiospores fall within 1 m of the cap: a field- and modeling-based study. Mycologia 103:1175. [4] Roper M, et al. (2010) Dispersal of fungal spores on a co- operatively generated wind. Proceedings of the National Academy of Sciences 107:17474–17479. [5] Roper M, Pepper RE, Brenner MP, Pringle A (2008) Ex- plosively launched spores of ascomycete fungi have drag- minimizing shapes. Proceedings of the National Academy of Sciences 105:20583–20588. [6] Fritz JA, Seminara A, Roper M, Pringle A, Brenner MP (2013) A natural O-ring optimizes the dispersal of fungal spores. J. R. Soc. Interface 10:20130187–20130187. [7] Money NP (1998) More g’s than the space shuttle: bal- listospore discharge. Mycologia pp 547–558. [8] Pringle A, Patek SN, Fischer M, Stolze J, Money NP (2005) The captured launch of a ballistospore. Mycologia 97:866. [9] Noblin X, Yang S, Dumais J (2009) Surface tension propulsion of fungal spores. Journal of Experimental Bi- ology 212:2835–2843. [10] Stolze-Rybczynski JL, et al. (2009) Adaptation of the spore discharge mechanism in the basidiomycota. PloS one 4:4163. [11] Buller AHR (1909) Researches on fungi (London, New York [etc.] Longmans, Green and co.). [12] Fischer MWF, Stolze-Rybczynski JL, Cui Y, Money NP (2010) How far and how fast can mushroom spores fly? Physical limits on ballistospore size and discharge dis- tance in the Basidiomycota. Fungal Biology 114:669–675. [13] Norros V, et al. (2014) Do small spores disperse further than large spores? http://dx.doi.org/10.1890/13-0877.1 95:1612–1621. [14] Aylor DE (1990) The role of intermittent wind in the dispersal of fungal pathogens. Annual Review of Phy- topathology 28:73–92. [15] Batchelor G (1967) Introduction to Fluid Dynamics (Cambridge University Press, Cambridge, U.K.). [16] Whitaker DL, Edwards J (2010) Sphagnum Moss Dis- perses Spores with Vortex Rings. Science 329:406. [17] Pedlosky J (1982) Geophysical fluid dynamics (Springer- Verlag, New York). [18] Mercier MJ, Ardekani AM, Allshouse MR, Doyle B, Pea- cock T (2014) Self-propulsion of immersed objects via natural convection. Physical Review Letters 112:204501. [19] Arora D (1986) Mushrooms Demystified, A Comprehen- sive Guide to the Fleshy Fungi (Springer Science & Busi- ness). [20] Husher J, et al. (1999) Evaporative cooling of mush- rooms. Mycologia 91:351–352. [21] Mahajan PV, Oliveira F, Macedo I (2008) Effect of tem- perature and humidity on the transpiration rate of the whole mushrooms. Journal of Food Engineering 84:281– 288. [22] Baehr H, Stephan K (2011) Heat and Mass Transfer 7 (Springer Berlin Heidelberg). [23] Huppert HE (2006) Gravity currents: a personal per- spective. Journal of Fluid Mechanics 554:299–322. [24] Fischer MWF, Money NP (2010) Why mushrooms form gills: efficiency of the lamellate morphology. Fungal Bi- ology 114:57–63. [25] Peay KG, Schubert MG, Nguyen NH, Bruns TD (2012) Measuring ectomycorrhizal fungal dispersal: macroeco- logical patterns driven by microscopic propagules. Molec- ular Ecology 21:4122–4136. [26] Sveen JK (2004) An introduction to matpiv v. 1.6. 1. Preprint series. Mechanics and Applied Mathematics http://urn. nb. no/URN: NBN: no-23418. [27] Roper M, Simonin A, Hickey PC, Leeder A, Glass NL (2013) Nuclear dynamics in a fungal chimera. Proceedings of the National Academy of Sciences 110:12875–12880. 8 FIG. 2: High rates of evaporation from the pileus cool the surrounding air. (A) A L. edodes pileus left in laboratory ambient conditions rapidly dries out. Time code in hh:mm format. Scale bar: 1cm. (B) Rates of water loss from mush- rooms greatly exceed plants. Mushrooms attached to intact mycelia (Mushrooms) lose water at a higher rate than living plants (Plants), and even agar hydrogels (Agar). Cut mush- rooms (Mush. (cut)) lose water at a higher rate than cut plant leaves (Plant (cut)). Spore liberation contributed negligibly to mass loss: when the gill surface was coated with petroleum jelly to prevent spore shedding (Mush. (cut/P)), measured mass loss was statistically identical to untreated mushrooms. (C) Evaporation cools the air beneath the pileus by several degrees ◦C, both for mushrooms stored in container to pre- vent external convection and for mushrooms maintained at laboratory ambient conditions, consistent with surface tem- perature measurements in [20]. Also shown: a convection control showing temperature variations in one of our experi- mental containers with no mushroom. FIG. 1: Mushroom spores readily disperse from pilei that are crowded close together or close to the ground. (A) Oyster mushrooms (Pleurotus ostreatus) grow crowded together in a stand. (B) Agaricus californicus mushrooms grow under a layer of plant litter on the UCLA campus. (C) Even when mushrooms are isolated from external airflows, an asymmetric tongue of spores is deposited far from the pileus. Scale bar: 5cm. (D) Extended asymmetric spore deposits for cultured (left data) as well as wild-collected circular pilei (right data). The parameter used to characterize the asymmetry of the spore deposit is described in the Materials and Methods. (E) Airflows carrying spores out from the gap can be directly observed using a laser light sheet. Shown: Lentinula edodes, scale bar: 1 cm. FIG. 3: Convective cooling produces a gravity current that disperses spores from beneath the pileus. Spore dispersal dis- tance (black data points) increases as (gap width)2 (black line), consistent with theory for asymmetrically shaped pilei (Equation (1)). 9 FIG. 4: Strong spore dispersal requires shape asymmetry or temperature differentials along the pileus. (A,B) Dispersal from a symmetrically cooled and shaped pileus is very weak (upper panel in (A), gray line in (B)). Imposing a temperature gradient across the pileus surface enhances dispersal (middle panel in (A), solid black line in (B)), as does asymmetric ar- rangement of the pileus (bottom panel in (A), dashed black line in (B)). Different forms of asymmetry produce different scalings for dispersal distance as a function of gap height. For temperature differentials, dmax ∼ h3 r, while for shape asymme- try dmax ∼ h2 r. Simulations are of a 4cm mushroom, with max surface ∆T = 3oC, colors show speed of convective flow, white lines are representative trajectories of spores with sedimenta- tion velocity 1mm/s, and yellow lines are flow streamlines. (C,D) A simple theory for the dispersal distance collapses data from different gap heights, spore sizes and amounts of asymmetry. (C) shows raw data for tilted pilei, with angles of tilt with the horizontal equal to 0o ((cid:74)), 10o ((cid:72)), 20o ((cid:73)), and pilei with left-right temperature differentials of 1oC ((cid:3)) and 1.5oC ((cid:13)). Two sedimentation velocities are shown: vs = 1 mm/s (black symbols) and 4 mm/s (gray symbols). (D) shows same data collapsed using the scaling derived in the main text. Asymmetric pilei, and pilei with temperature dif- ferentials lie on different lines, but in both cases dispersal distance is proportional to [h2∆T ]hr/vs (gray lines). (E,F) We test the scalings for real mushrooms using Digital PIV to measure spore velocities. In (E) colors give spore velocity in m/s, scale bar: 1cm). In (F), the mean spore velocity at the beginning of the gravity current is proportional to gap width h (black line), consistent with Equation (1) for shape-induced asymmetry. 10 FIG. 5: Nearby boundaries enhance convective spore disper- sal. (A) When a mushroom is surrounded by a circular bar- rier, spores can disperse over the barrier. Here we shifted the mushroom and barrier at the end of the experiment, so that the spore deposit can be seen more clearly by the contrast with the black circle of clean transparency beneath the bar- rier. In general spores that travel over the barrier were uni- formly dispersed over the entire footprint of the box that en- closed the experiment, regardless of their horizontal spreading range. (B) Spores were able to cross a barrier if the height of the barrier was less than the horizontal range of spores (black line). Colors in the scatter plot correspond to a measure of barrier crossing rate (see Materials and Methods), in partic- ular dark blue points correspond to mushrooms whose spores did not cross the barrier. (C) Spores climbing the barrier are observed being swept into the eddy of warm air that feeds the gravity current, potentially explaining their enhanced dis- persal after crossing the barrier (shown: spore current from a L. edodes pileus). Scale bars: 2 cm. (D) Spores climb- ing an inclined barrier (black curve, numerical simulations) travel further than along horizontal surfaces (gray curve gives 1/ cos θ× the horizontal range of the spores). Supplementary Information for: “Mushroom spore dispersal by convectively-driven winds” This document contains the following Supplementary Figures associated with the paper: “Mushroom spore dispersal by convectively-driven winds”, S1 Demonstration that spore deposition patterns rotate with the mushroom. S2 Spore dispersal distance does not depend on the pileus diameter or on the rate of spore production. S3 Numerical simulations show that gravity currents can carry spores up barriers. 11 FIG. S1: To check that the airflows carrying spores out from under the pileus were created by the pileus, and not by external temperature gradients, we performed experiments in which the transparency was replaced in the middle of the experiment and the mushroom was rotated either by 90o or 180o. In all cases we saw that the pattern of spore deposition rotated with the mushroom. Two representative experiments are shown here. The orange shapes show the approximate shape of the pileus, and arrows show orientation of arbitrary reference points on the pileus. (A-B) Deposition from a P. ostreatus mushroom. (A) Deposition after 2 hours. (B) The mushroom was rotated by 180o, a new transparency added, and the experiment continued for another 2 hours. (C-D) Deposition from a L. edodes mushroom. (C) Deposition after 2 hours. (D) The mushroom was rotated by 90o, a new transparency was added, and the experiment continued for another 2 hours. FIG. S2: Distance spores dispersed from under the pileus does not depend on the pileus diameter or on the rate of spore production. Here data from 30 P. ostreatus mushrooms are shown (a subset of the experiments from main text Fig. 3). (A) Dispersal distance does not correlate with pileus diameter (R2 = 0.07). (B) Dispersal distance does not correlate with the rate of spore production (q, number of spores released per cm2 of pileus, per s, R2 = 0.17). 12 FIG. S3: Cold air leaving the gap beneath the pileus is replaced by pulling warm air to the pileus. In the presence of an external wall, numerical simulations (see main text: Materials and Methods), cold outflow and warm inflow can connect to form a closed convective eddy. Our simulations show that spores in the cold gravity current may be drawn up the wall by the closed eddy. Here, gray curves show the streamlines of the flow created by the cooled pileus. Black lines show simulated spore trajectories. The external wall is made up of two slanted surfaces on the right hand side of the image.
1210.6250
1
1210
2012-10-23T14:33:58
Theory and Monte Carlo simulations for the stretching of flexible and semi-flexible single polymer chains under external fields
[ "physics.bio-ph", "cond-mat.soft", "math-ph", "math-ph" ]
Recent developments of microscopic mechanical experiments allow the manipulation of individual polymer molecules in two main ways: \textit{uniform} stretching by external forces and \textit{non-uniform} stretching by external fields. Many results can be thereby obtained for specific kinds of polymers and specific geometries. In this work we describe the non-uniform stretching of a single, non-branched polymer molecule by an external field (e.g. fluid in uniform motion, or uniform electric field) by a universal physical framework which leads to general conclusions on different types of polymers. We derive analytical results both for the freely-jointed chain and the worm-like chain models based on classical statistical mechanics. Moreover, we provide a Monte Carlo numerical analysis of the mechanical properties of flexible and semi-flexible polymers anchored at one end. The simulations confirm the analytical achievements, and moreover allow to study the situations where the theory can not provide explicit and useful results. In all cases we evaluate the average conformation of the polymer and its fluctuation statistics as a function of the chain length, bending rigidity and field strength.
physics.bio-ph
physics
Theory and Monte Carlo simulations for the stretching of flexible and semi-flexible single polymer chains under external fields Fabio Manca,1 Stefano Giordano,2, 3, a) Pier Luca Palla,2, 4 Fabrizio Cleri,2, 4 and Luciano Colombo1 1)Department of Physics, University of Cagliari, 09042 Monserrato, Italy 2)Institute of Electronics, Microelectronics and Nanotechnology (UMR CNRS 8520), 59652 Villeneuve d'Ascq, France 3)International Associated Laboratory LEMAC, ECLille, 59652 Villeneuve d'Ascq, France 4)University of Lille I, 59652 Villeneuve d'Ascq, France (Dated: 27 March 2021) Recent developments of microscopic mechanical experiments allow the manipulation of individual polymer molecules in two main ways: uniform stretching by external forces and non-uniform stretching by external fields. Many results can be thereby obtained for specific kinds of polymers and specific geometries. In this work we describe the non-uniform stretching of a single, non-branched polymer molecule by an external field (e.g. fluid in uniform motion, or uniform electric field) by a universal physical framework which leads to general conclusions on different types of polymers. We derive analytical results both for the freely-jointed chain and the worm-like chain models based on classical statistical mechanics. Moreover, we provide a Monte Carlo numerical analysis of the mechanical properties of flexible and semi-flexible polymers anchored at one end. The simulations confirm the analytical achievements, and moreover allow to study the situations where the theory can not provide explicit and useful results. In all cases we evaluate the average conformation of the polymer and its fluctuation statistics as a function of the chain length, bending rigidity and field strength. I. INTRODUCTION Modern methods for stretching single molecules pro- vide a valuable insight about the response of polymers to external forces. The interest on single molecules load- ing encouraged new research and technological develop- ments on related mechanical experiments. Typically, me- chanical methods allow the manipulation of a polymer molecule in two ways: the stretching of the chain by the direct action of an external force or by the application of an external field. If we consider homogeneous polymers (with all monomers described by the same effective elas- tic stiffness), then we obtain a uniform strain with the external force and a non-uniform strain with the applied field. To exert an external force on a polymer fixed at one end, laser optical tweezers (LOTs)1, magnetic tweezers (MTs)2 or atomic force microscope (AFM)3 can be used. Many experiments have been performed over a wide class of polymers with biological relevance, such as the nucleic acids (DNA, RNA)4, allowing the stretching of the en- tire molecule and providing the reading and the mapping of genetic information along the chain.5,6 Furthermore, it has been possible to describe the elastic behaviour of single polymers consisting of domains which may exhibit transitions between different stable states.7 -- 9 Other in- vestigations performed on double-stranded DNA deter- mined the extension of the polymer as a function of the applied force10, providing results in very good agreement with the Worm-Like Chain (WLC) model11 -- 13 and the Freely-Jointed Chain (FJC) model.13,14 a)Electronic mail: [email protected] Alternatively, it is possible to manipulate single molecules by an external field. In this case the external field acts on the molecules from a distance or, in other words, without a defined contact point for applying the traction. A non-uniform stretching performed by an ex- ternal field can be induced either via a hydrodynamic (or electrohydrodynamic) flow field15 -- 17 or via an electric (or magnetic) field.18 -- 20 One experimental advantage of us- ing flow fields is that the liquid surrounding the tethered molecule can be easily replaced; this is indeed an im- portant feature for many single-molecules studies of en- zymes which require varying buffer conditions.21 The flow field technique was extensively applied in single-molecule study of DNA elasticity11 as well as to characterize the rheological properties of individual DNA molecules.22 -- 24 The use of an electric field has been adopted for driving the alignment of DNA on a solid surface for applications such as gene mapping and restriction analysis.18 Finally, magnetic fields have been used to apply torsional stress to individual DNA molecules.19,20 In order to understand the response of polymers to ex- ternal fields and to study their statistics, some theoretical models have been proposed. These models are typically based on the FJC and WLC schemes, generalized with the inclusion of the given applied field. Some studies have shown that in a weak external field the persistence length along the field direction is increased, while it is decreased in the perpendicular direction; moreover, as the external field becomes stronger, the effective persistence length grows exponentially with the field strength.25 -- 27 Other investigations under a constant velocity flow have shown that a flexible polymer displays three types of confor- mation: unperturbed at low velocity; "trumpet" shaped when partially stretched; "stem and flowers" shaped, 1 2 1 0 2 t c O 3 2 ] h p - o i b . s c i s y h p [ 1 v 0 5 2 6 . 0 1 2 1 : v i X r a with a completely stretched portion (the stem) and a se- ries of blobs (the flowers), at larger loading.28 -- 30 Polymer models have been studied in elongational flows to analyze the coil stretching and chain retraction as a function of polymer and flow parameters, finding good agreement with experimental data.31,32 Conformational properties of semiflexible polymer chains in uniform force field were also studied for two-dimensional models.33 In spite of all these relevant efforts, it is yet a challenge to base on one same unified theoretical framework and understanding of all aspects of polymer mechanics in an external field. k ~r0 FIXED ~g1 ~r1 k ~r2 ~g2 ~rK ~gK k ~gN −1 ~rN −1 k ~f ~gN ~rN Building on our previous studies,9,13 in this paper we study the conformational and mechanical properties of flexible and semi-flexible non-branched polymer model chains tethered at one end and immersed in an external force field. This situation is useful to describe almost two physical conditions of interest: a polymer chain immersed in a fluid in a uniform motion (our model is valid only when the action of the fluid motion can be described by a distribution of given forces applied to all monomers) and an arbitrarily charged chain inserted in a uniform electric field. Our theoretical approach is twofold, since we adopt both analytical (statistical mechanics34,35) and numer- ical techniques (Monte Carlo simulations36,37). While the analytical approach is useful to obtain the explicit partition function in some specific cases, Monte Carlo simulations are crucial to study more generic cases, in- accessible to analytical treatments. In particular, while we develop our theoretical framework starting from the more tractable FJC model, we take full profit from our MC simulations to extend our study also to the WLC model. The structure of the paper is the following. In Section II we introduce the mathematical formalism adopted and we derive a generic form of the partition function in ℜd for a generalized FJC model where the extensibility of the bonds is taken into account. In the Section III we find the two specific forms of the partition function for the 2D- and the 3D-case for the pure FJC polymer with non extensible bonds. Moreover, we obtain in both cases the variance and the covariance among the positions of the monomers. In the Section IV we present the gen- eralization of previous results to the semi-flexible WLC model. We present two closed-forms approximations for the 2D- and the 3D-case and the comparisons with MC simulations. In section V we analyze the behavior of a chain in an external field to which also an external force is applied at the end of the chain. The case with the force not aligned with the field is particularly interesting and shows the power of the MC method. Finally, in Section VI some conclusions are drawn. 2 FIG. 1. (color online) A polymer chain in an external field. The first monomer is clamped at position ~r0 while the oth- ers are free to fluctuate. Each monomer is subjected to an external force ~gK (different in strength and direction for any K): all these forces mimic an external field. Another external force, playing the role of a main pulling load, ~f , is applied to the last monomer at the position ~rN . II. GENERAL THEORETICAL FRAMEWORK As previously discussed, the polymer models most used in literature are the FJC and the WLC. As argued in Ref.38, for weak tension and weak external field, it is ac- ceptable to model the polymer as a FJC model. This model breaks down only when the curvature of the con- formation is very large because it ignores the consequent great bending energy. Since we will look upon this prob- lem in the end of this work, we now give way to the case of a FJC. In particular we consider a FJC with two additional hypothesis. Firstly we consider the possible extensibility of the bonds of the chain through a stan- dard quadratic potential characterized by a given equi- librium length: such an extension mimics the possible stretching of the chemical bond between two adjacent monomers. If necessary, the extensibility of the bonds, here described by linear springs, can be easily extended to more complex, nonlinear springs.39 Moreover, we take into account a series of arbitrary forces applied to each monomer: these actions mimic the effects of an external physical field applied to the system. In addition, we con- template the presence of an arbitrary force applied to the terminal monomer of the chain. All calculations will be performed in ℜd and we will specialize the results both in the 2D-case and in the 3D-case when needed. The idea is to write the complete form of the Hamiltonian of the system and to build up the corresponding statistical mechanics.13 The starting point is therefore the calcula- tion of the classical partition function. In fact, when this quantity is determined, it is possible to obtain the force- extension curve (the equation of state) through simple derivations. Let us consider a non-branched linear polymer with N monomers (see Fig. 1) at positions defined by ~r1, ..., ~rN ∈ ℜd (for considering d = 2 or d = 3 according to the specific problem of interest). To each monomer a given external force is applied and named ~g1, ..., ~gN . Another external force, playing the role of main pulling load, ~f , is applied to the last monomer at the position ~rN . While the chain is clamped at position ~r0, the monomers are free to fluctuate. The Hamiltonian of the system is therefore given by H = (~rK − ~rK−1 − l)2 N ~pi · ~pi 2m 1 2 k + XK=1 ~gK · ~rK − ~f · ~rN N Xi=1 XK=1 − N (1) where ~pi are the linear momenta, m the mass of the monomers, k the spring constant of the inter-monomer interaction, and l the equilibrium length of the monomer- monomer bond. We search for the partition function of the system defined as: exp(cid:18)− H kBT(cid:19) d~r1...d~rN d~p1...d~pN (2) Zd = cZℜd ...Zℜd {z } 2N −times where c is a multiplicative constant which takes into ac- count the number of microstates. As well known, the kinetic part can be straightforwardly integrated and it yields a further non-influencing multiplicative constant; then we can write the partition function as an integral over the positional space only. This integral can be easily handled through the standard change of variable partition function becomes N −times ...Zℜd Zd = cZℜd {z } × exp" 1 × exp" 1 kBT kBT exp"− k 2kBT N XK=1(cid:16)~ξK − l(cid:17)2# (5) ~ξJ# N K XK=1 XJ=1 ~gK · ~ξK# d~ξ1...d~ξN XK=1 ~f · N Inverting the two summation symbols N XK=1 ~gK · K XJ=1 ~ξJ = N XK=1 ~ξK · ~gi N Xi=K Zd = c N YK=1Zℜd e−a(~ξ−l)2 ~VK ·~ξd~ξ e we obtain where a = k 2kBT > 0 ~VK = 1 kBT ~f + ~gi! N Xi=K (6) (7) (8) (9) It exists a deep conceptual connection between the last integral for the partition function and the theory of the d-dimensional Fourier transforms. The Fourier integral of an arbitrary function f (~ξ) is defined as F (~ω) =Zℜd f (~ξ)e−i~ω·~ξd~ξ (10) ~ξ1 = ~r1 − ~r0 ~ξ2 = ~r2 − ~r1 ... ~ξN = ~rN − ~rN −1 with inverse transform given by (3) f (~ξ) = 1 (2π)d Zℜd F (~ω)ei~ω·~ξd~ω (11) If we consider   having the Jacobian determinant J = (cid:12)(cid:12)(cid:12) We consider the terminal ~r0 of the chain fixed in the origin of axes, i.e. ~r0 = ~0. So, we cast the positions ~ri in terms of the variables ~ξJ as follows ∂(~ξ1...~ξN )(cid:12)(cid:12)(cid:12) = 1. ∂(~r1...~rN ) ~r1 = ~ξ1 + ~r0 = ~ξ1 ~r2 = ~ξ2 + ~r1 = ~ξ2 + ~ξ1 ... ~rN = ~ξN + ~ξN −1 + ... + ~ξ1   (4) By setting the general solution as ~ri = Pi K=1 ~ξK, the 3 f (~ξ) = e−a(~ξ−l)2 (12) it is easy to realize that the integral in Eq.(7) is the Fourier transform of f (~ξ) calculated for ~ω = i~VK, i.e. Zd = c N YK=1 F (i~VK ) (13) with a e ~VK defined respectively in Eq.(8) and Eq.(9). It is important to remark that the function in Eq.(12) has a spherical symmetry (i.e. it depends only on the length of the vector ~ξ) and, therefore, also its Fourier transform F (~ω) exhibits the spherical symmetry, depending only on the quantity ~ω in the transformed domain. In fact, for such spherically-symmetric functions it holds that: if f (~ξ) = f (~ξ) then F (~ω) = F (~ω). Furthermore, we have that F (Ω) =Z +∞ Ω (cid:19) d 2πρf (ρ)(cid:18) 2πρ 2 −1(ρΩ)dρ (14) 2 −1 J d 0 for d = 2n (even), and 2 2 0 (15) j d−3 (ρΩ)dρ Ω (cid:19) d−3 4πρ2f (ρ)(cid:18) 2πρ F (Ω) =Z +∞ for d = 2n + 1 (odd), where ρ = ~ξ and Ω = ~ω.40 Here Jν(z) and jν (z) are the cylindrical and spherical Bessel functions of the first kind respectively, correlated by the standard relation jν(z) =p π (z).41,42 In our calculations we have to set ~ω = i~VK and, therefore, we obtain Ω = i~VK. Moreover, when the argument of Jν (z) and jν(z) is supposed imaginary we obtain the modified Bessel functions of the first kind41,42 2z Jν+ 1 2 Iν (z) = (i)−ν Jν(iz) iν(z) = (i)−ν jν(iz) (16) For example we have the explicit expression j0(z) = sin z z and i0(z) = sinh z z while, on the contrary, I0(z) and J0(z) cannot be written in closed form. So, for d even we even- tually obtain III. FREELY-JOINTED CHAIN MODEL UNDER EXTERNAL FIELD A. Average values of positions In the previous section we obtained the general ex- pression of the partition function for the case where the extensibility of the bonds is taken into account. This is described by the parameter k, which characterizes the elastic bond between adjacent monomers. In the present Section we want to study the effects of an arbitrary dis- tribution of forces on a pure freely jointed chain model (FJC). Therefore we need to obtain the specific form of the partition function in the case of rigid bonds of fixed length l. From the mathematical point of view it means that we will consider k → ∞, a condition rep- resenting a inextensible spring. Because of the relation p α = δ(x) when α → ∞ we may determine the limit of Eq.(19) and Eq.(20) for a → ∞ (i.e. for k → ∞, FJC limit). Since the arbitrariness of the constant c, we may consider in Eqs. (19) and (20) a multiplicative constant term (p a π )N . Then, by using the translated propertyp a → δ(ρ − l) for a → ∞ we perform YK=1 all the integrals thereby obtaining π e−a(ρ−l)2 1 ~VK (l~VK) π e−αx2 Zd = c d even I d−2 (21) d−2 N 2 2 F (i~VK) 2π =Z +∞ 0 ρ e−a(ρ−l)2 2πρ ~VK! d−2 2 I d−2 2 (ρ~VK)dρ (17) and, on the other hand, for d odd we have Zd = c N YK=1 1 ~VK i d−3 2 (l~VK) d−3 2 d odd (22) In particular, for d = 2 we have 4π F (i~VK) =Z +∞ ρ2 e−a(ρ−l)2 2πρ ~VK! d−3 (ρ~VK)dρ (18) Finally, by using Eq.(13), the partition function is given by i d−3 0 2 2 Zd = c N YK=1Z +∞ 0 for d even, and ρ e−a(ρ−l)2 ρ 2 ~VK! d−2 I d−2 2 (ρ~VK)dρ (19) N YK=1Z +∞ 0 ρ2 e−a(ρ−l)2 ρ 2 ~VK! d−3 2 i d−3 Zd = c (ρ~VK)dρ (20) for d odd, where a and ~VK are given in Eqs.(8) and (9). In the framework of statistical mechanics, the knowledge of the partition function allows to determine all needed expected values describing the statistics of the chain (i.e., average values of the positions, variances of the positions and so on). while for d = 3 we obtain (23) (24) Z2 = c Z3 = c N YK=1 N YK=1 N ~f + ! kBT (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ~gi(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) I0 l Xi=K i=K ~gi(cid:12)(cid:12)(cid:12)(cid:17) kB T (cid:12)(cid:12)(cid:12) sinh(cid:16) l ~f +PN i=K ~gi(cid:12)(cid:12)(cid:12) kB T (cid:12)(cid:12)(cid:12) ~f +PN l All the expressions given in Eqs.(21), (22), (23), (24) can be summarized in the general form Zd = c N YK=1 f (~VK) (25) with a suitable function f (x). By using this expression of the partition function we can find the average position of the i-th monomer of the chain; indeed, from the definition of the Hamiltonian in Eq.(1) we state that ~ri = − ∂H and, therefore, we get ∂~gi h~rii = kBT ∂ ∂~gi ln Zd (26) 4 which represents the shape of the polymer chain under the effects of the external field ~gi and the applied force ~f . Now we can substitute Eq.(25) into Eq.(26), obtaining In 2D we have f (x) = I0(lx) and therefore we obtain h~rii = f (x) i ~VK ~VK(cid:20) 1 XK=1 kB T (cid:12)(cid:12)(cid:12) I1(cid:16) l kB T (cid:12)(cid:12)(cid:12) I0(cid:16) l ~f +PN ~f +PN h~rii = l i XK=1 ∂f (x) ∂x (cid:21)x=~VK (27) J=K ~gJ(cid:12)(cid:12)(cid:12)(cid:17) J=K ~gJ(cid:12)(cid:12)(cid:12)(cid:17) ~f +PN ~f +PN J=K ~gJ J=K ~gJ(cid:12)(cid:12)(cid:12)(28) (cid:12)(cid:12)(cid:12) For such a 2D case, by applying Eq.(28), the average val- ues of the longitudinal component of the positions have been calculated and are plotted in Fig.2 as a function of the chain length N and the field strength g. We have considered only the action of an external uniform field with ~gJ = ~g and amplitude g. Although this case lends itself to a full analytical so- lution, numerical simulations were also performed by us- ing a conventional implementation of the Metropolis ver- sion of the Monte Carlo algorithm.36 The initial state of the chain is defined by a set of randomly chosen posi- tions. The displacement extent of each step governs the efficiency of the configurational space sampling. There- fore, we analysed several runs in order to optimize its value.43,44 The perfect agreement between the theory and the MC simulations provides a strict check of the numer- ical procedure, to be used in the foregoing. On the other hand, in 3D we have f (x) = sinh(lx) , lx leading to h~rii = l i XK=1 kBT (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) L l ~f + N XJ=K ~gJ(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ! ~f +PN (cid:12)(cid:12)(cid:12) ~f +PN J=K ~gJ J=K ~gJ(cid:12)(cid:12)(cid:12)(29) where L(x) = coth x − 1 x is the Langevin function. By using Eq.(29), as before, it is possible to plot the aver- age values of the longitudinal component of the positions for the 3D case (Fig.3). Also in this case we adopted a uniform field g and the good agreement with the MC simulations is evident. As particular case, if there is only the force ~f applied to the system we obtain the standard scalar force-extension curves linking r = h~rNi with f = ~f. In 2D we have in agreement with recent results,45 while in 3D we obtain r lN r lN = kB T(cid:17) I1(cid:16) lf I0(cid:16) lf kB T(cid:17) kBT(cid:19) = L(cid:18) lf (30) (31) which is a classical result.13,46 The simple results in Eqs.(30) and (31) have been used to obtain the limiting 5 50 40 30 20 10 0 0 25 20 15 10 5 0 0 i/l l , ri h i/l l , ri h 2D−FJC N 0.5 i/N 1 2D−FJC g 0.5 i/N 1 FIG. 2. (color online) Average values of the longitudinal com- ponent of the positions induced by the external field for the 2D FJC case. The red solid lines correspond to the analyt- ical results Eqs.(28) and (32), MC results are superimposed in black circles. Top panel: each curve corresponds to dif- ferent chain lengths N = 10, 20, 30, 40, 50 for a fixed value gl/(kBT ) = 1 (e.g., corresponding to l = 1nm, g = 4pN at T = 293K). Bottom panel: each curve corresponds to the different values gl/(kBT ) = 0.1, 0.25, 0.5, 1, 2, 10 for a fixed chain length N = 20. behaviors under low (f → 0) and high (f → ∞) values of the applied force, as shown in Table I. Building on such first results we now focus on some particular interesting approximations. More specifically, it can be interesting to find approximate results for the case of a homogeneous field and no end-force, ~f = 0 and ~gJ = ~g for any J. In this case we search for the scalar relation between r = h~rNi and g = ~g. In the 2D case, from Eq.(28), we have r lN = ≃ = N 1 N Xk=1 N Z N 1 0 1 N 1 lg kB T kB T (N − k + 1)(cid:17) I1(cid:16) lg kB T (N − k + 1)(cid:17) I0(cid:16) lg kB T (N − x + 1)(cid:17) I1(cid:16) lg kB T (N − x + 1)(cid:17) dx I0(cid:16) lg kB T (N + 1)(cid:17) I0(cid:16) lg kB T(cid:17) I0(cid:16) lg log (32) 50 40 30 20 10 0 0 20 15 10 5 0 0 i/l l , ri h i/l l , ri h 3D−FJC N 0.5 i/N 1 3D−FJC g 0.5 i/N 1 FIG. 3. (color online) Average values of the longitudinal com- ponent of the positions induced by the external field for the 3D FJC case. The red solid lines correspond to the analyt- ical results Eqs.(29) and (33), MC results are superimposed in black circles. Top panel: each curve corresponds to dif- ferent chain lengths N = 10, 20, 30, 40, 50 for a fixed value gl/(kBT ) = 1. Bottom panel: each curve corresponds to the different values gl/(kBT ) = 0.1, 0.25, 0.5, 1, 2, 10 for a fixed chain length N = 20. B. Covariances and variances of positions In this Section, we search for the covariance among the positions of the monomers. It is important to evalu- ate such a quantity in order to estimate the variance of a given position (measuring the width of the probabil- ity density around its average value) and the correlation among different monomer positions (measuring the per- sistence of some geometrical features along the chain). In order to do this, we identify the α-th component of the i-th monomer as riα. The covariance of the generic monomer simply defined as (it represent the expectation value of the second order): Cov(riα, rJβ) = h(riα − hriαi)(rJβ − hrJβi)i = hriαrJβi − hriαihrJβi (34) Taking the derivative of the partition function with re- spect to the α and the β components of the force vectors ~gi and ~gJ we can solve the problem as follows. We con- sider the standard expression for the partition function and we can elaborate the following expression hriαrJβi = (kBT )2(cid:18) ∂ ln Zd ∂giα ∂ ln Zd ∂gJβ + ∂2 ln Zd ∂giα∂gJβ(cid:19)(35) or, equivalently, by introducing Eq.(26) hriαrJβi = hriαihrJβi + kBT ∂ ∂gJβ hriαi (36) but we can simply determine that ∂ ∂gJβ hriαi = ∂ ∂gJβ i XK=1 ~VK (cid:20) 1 ~VK · ~eα f (x) ∂f (x) ∂x (cid:21)x=~VK (37) On the other hand, for the 3D case we obtain where we have defined the unit vector ~eα as the basis of the orthonormal reference frame. Being r lN = ≃ = N 1 N L(cid:18) l XK=1 0 L(cid:18) l N Z N 1 1 N 1 lg kB T log kBT kBT (N − k + 1)(cid:19) (N − x + 1)(cid:19) dx e2 lg kB T (N +1) − 1 kB T − 1(cid:17) − 1 (N + 1)(cid:16)e2 lg and (33) ~VK · ~eα = 1 kBT fα + giα! N Xi=K (38) ∂~VK ∂gJβ = 1 kBT ~VK · ~eβ VK N Xq=K δJq (39) after long but straightforward calculations we obtain We have usefully exploited the fact that, for large N , the sums can be approximately substituted with the corre- sponding integrals, which are easier to be handled. The closed-form expressions given in Eqs.(32) and (33) are very useful to obtain the limiting behaviors of the poly- mer under low (g → 0) and high (g → ∞) values of the applied field, as shown in Table I. Moreover, we have verified the validity of Eqs.(32) and (33) through a series of comparisons with MC results (see Fig.6 in the next Section for details). 6 (40) min{i,J} 1 kBT ∂ ∂gJβ hriαi = XK=1 ×(δαβf ′(~VK) + f ′′(~VK) − VKαf ′(~VK) ~VK2 − VKα VKβ ~VKf (~VK) VKαVKβ ~VK f ′(~VK)2 f (~VK) VKβ ~VK) Ordering the terms we finally obtain the important result Cov(riα, rJβ ) = + ×(f ′′(~VK) − min{i,J} min{i,J} XK=1 XK=1 f ′(~VK) f (~VK) δαβ ~VK VKαVKβ ~VK2f (~VK) f ′(~VK) ~VK (41) f (~VK) ) f ′(~VK)2 − It represents the final form of the covariance between two different components of the positions of two different monomers. If we look at the variance of a single component of a single position (i = J, α = β) we have the simpler result σ2 iα = i i V 2 f ′(~VK) XK=1 ~VKf (~VK) ×(f ′′(~VK) − + Kα XK=1 ~VK2f (~VK) f (~VK) ) f ′(~VK) f ′(~VK)2 − ~VK (42) lx In order to use the previous expressions we have to specify the function f and its derivatives for the two- dimensional and the three-dimensional case. In the 2D case we have f (x) = I0(lx), f ′(x) = lI1(lx) and f ′′(x) = l2 2 [I0(lx) + I2(lx)]. On the other hand, for the 3D case we have f (x) = sinh(lx) , f ′(x)/f (x) = lL(lx) and f ′′(x)/f (x) = l2 − 2lL(lx)/x. This completes the determination of the covariance. We report in Fig.4 and Fig.5 the longitudinal and transversal component of the variance as a function of the chain length and the field strength for the 3D case (with f = 0). The 2D case is very similar and it has not been reported here for sake of brevity. We can observe some interesting trends: the longitudinal variance of the posi- tion is a decreasing function of the number of polymers N while the transversal one is a increasing function (with a fixed amplitude of the external field g). Moreover, both variances are rapidly increasing along the chain, assum- ing the largest value in the last free monomer, which is more subject to strong fluctuations. It interesting to ob- serve that the variance (both longitudinal and transversal components) is a linear function of the position i along the chain (it linearly intensifies along the chain itself) with a simple force f applied at the free end: conversely, with a uniform field g, the distribution of forces generates a strongly non-linear intensification of the variances mov- ing towards the free end-terminal. So, from the point of view of the variances, the application of a field or the ap- plication of a single force generates completely different responses. In Fig.5 we can also observe that the vari- ances are decreasing functions of the strength of the field (both for the longitudinal and transversal components); in fact, the intensity of the fields tends to reduce the fluc- tuations of the chain, increasing, at the same time, the tension within the bonds. 7 3D−FJC 1 0.8 0.6 0.4 0.2 2 /l l , 2 i σ N 0 0 4 3 2 1 0 0 2 /l t , 2 i σ 0.5 i/N 1 3D−FJC N 0.5 i/N 1 FIG. 4. (color online) Longitudinal (top panel) and transver- sal (bottom panel) component of the variance of positions for the 3D FJC case. The red solid lines correspond to the ana- lytical result Eq.(42), MC results are superimposed in black circles. Each curve corresponds to different chain lengths N = 10, 20, 30, 40, 50 for a fixed value of the external field defined by gl/(kBT ) = 1. IV. WORM-LIKE CHAIN MODEL UNDER EXTERNAL FIELD In previous Sections we treated systems described by the FJC model, characterized by the complete flexibil- ity of the chain and, therefore, by the absence of any bending contribution to the total energy. Neverthe- less, in many polymer chains, especially of biological ori- gin, the specific flexibility (described by the so-called persistence length47) has a relevant role in several bio- mechanical processes. In order to take into consideration these important features, with relevant applications to bio-molecules and bio-structures, in this Section we in- troduce the semi-flexible polymer chain characterized by a given bending energy added to the previous Hamilto- nian H = N Xi=1 + 1 2 κ N 1 2 ~pi · ~pi 2m N −1 k + XK=1 Xi=1 (cid:0)~ti+1 − ~ti(cid:1)2 (k~rK − ~rK−1k − l)2 XK=1 ~gK · ~rK − ~f · ~rN − N (43) where κ is the bending stiffness, k is the stretching mod- ulus and ~ti = (~ri+1 − ~ri)/k~ri+1 − ~rik is the unit vector 8 6 4 2 0 0 8 6 4 2 0 0 2 /l l , 2 i σ 2 /l t , 2 i σ 3D−FJC g F (f l/(kBT )). When ~f = 0 and ~gJ = ~g for any J we search for the 2D scalar relation between r and g = ~g. As discussed in a previous section (see Eqs.(32) and (33)), we can write 0.5 i/N 1 3D−FJC g r lN = ≃ = N 1 kBT 1 N F(cid:18) lg Xk=1 k=0 F(cid:18) lg N Z N kB T Z lg 1 N 1 lg kB T lg kBT (N − k + 1)(cid:19) (N − x + 1)(cid:19) dx F (y) dy kB T (N +1) (45) lg where we have defined the change of variable y = kB T (N − x + 1). We adopt now a second change of vari- able through the relation z = F (y) or y = F −1(z); it leads to r lN = = 1 N 1 N 0.5 i/N 1 kB T (N +1)(cid:17) 1 lg kB T Z F(cid:16) lg kB T (cid:17) F(cid:16) lg l Lp 1 lg kB T z F −1 (z) dz dz (46) FIG. 5. (color online) Longitudinal (top panel) and transver- sal (bottom panel) component of the variance of positions for the 3D FJC case. The red solid lines correspond to the analyt- ical result Eq.(42), MC results are superimposed in black cir- cles. Each curve corresponds to different values of the external field amplitude defined by gl/(kBT ) = 0.1, 0.25, 0.5, 1, 2, 10 for a fixed chain length N = 20. collinear with the i-th bond (see Ref.13 for details). In particular we take into consideration the classical WLC model, describing an inextensible semi-flexible chain: it means that the spring constant k is set to a very large value (ideally k → ∞) so that the bond lengths remain fixed at the value l. It is well known that it is not possible to calculate the partition functions in closed form for the WLC polymers. Nevertheless, some standard approxi- mations exist for such cases leading to simple expres- sions for the force-extension curves when a single force f is applied to one end of the chain. In the following, start- ing from these results, we search for the force-extension curves when the polymers is stretched through a constant field g. We start with the result for the 2D-WLC with an ap- plied force f : the approximated force extension curve is given by48 f l kBT = l Lp(cid:20) 1 16(1 − ζ)2 − 1 16 + 7 8 ζ(cid:21) (44) where ζ = r/(lN ) is the dimensionless elongation and Lp = lκ/(kBT ) is the persistence length . We suppose that such a constitutive equation is invertible through the function F , leading to the expression ζ = r/(lN ) = ×(cid:20) 7 16 z2 − 1 8(1 − z) + kB T (N +1)(cid:17) 1 16(1 − z)2(cid:21)F(cid:16) lg F(cid:16) lg kB T (cid:17) where we used the notation [h(z)]b a = h(b) − h(a). This result represents (although in implicit form) the approx- imated force-extension curve for the 2D-WLC under ex- ternal fields. To evaluate Eq.(46) we need to know the inverse function F (·), a task that can be performed nu- merically. Similarly, we may consider the standard 3D-WLC model with an applied force f ; the classical Marko-Siggia result12 is f l kBT = l Lp(cid:20) 1 4(1 − ζ)2 − 1 4 + ζ(cid:21) (47) where, as before, ζ = r/(lN ) is the dimensionless elon- gation and Lp = lκ/(kBT ) is the persistence length. We suppose again that such constitutive equation is invert- ible through the function G, leading to the expression ζ = r/(lN ) = G(f l/(kBT )). When ~f = 0 and ~gJ = ~g for any J we search for the 3D scalar relation between r and g = ~g. By repeating the previous procedure, we can write r lN = = 1 N 1 N 1 lg kB T kB T (N +1)(cid:17) 1 lg kB T Z G(cid:16) lg kB T (cid:17) G(cid:16) lg l Lp z G−1 (z) dz dz (48) ×(cid:20) 1 2 z2 − 1 2(1 − z) + 8 kB T (N +1)(cid:17) 1 4(1 − z)2(cid:21)G(cid:16) lg G(cid:16) lg kB T (cid:17) 1 0.8 0.6 l r N 0.4 0.2 0 0 1 1 0.8 0.6 l r N 0.4 0.2 0 0 1 5 5 1 0.8 0.6 l r N 0.4 0.2 0 0 1 1 0.8 0.6 l r N 0.4 0.2 0 0 1 2D FJC external force f=g FJC external force f=Ng FJC external field g MC FJC external field g 2 lg 4 3 kB T 3D FJC external force f=g FJC external force f=Ng FJC external field g MC FJC external field g 2 lg 3 4 kB T 5 5 2D WLC external force f=g WLC external force f=Ng WLC external field g MC WLC external field g 2 lg 3 4 kB T 3D WLC external force f=g WLC external force f=Ng WLC external field g MC WLC external field g 2 lg 3 4 kB T FIG. 6. (color online) Force-extension curves of a FJC poly- mer in an external field (or external force) with N=20. The red line corresponds to the approximated expressions given in Eqs.(32) and Eqs.(33) while the black circles have been obtained through MC simulations. The 2D (Eq.(30)) and 3D (Eq.(31)) FJC expressions (without an external field) are plotted for comparison with f = g and f = N g. which represents the implicit form of the approximated force-extension curve for the 3D-WLC under external fields. It is interesting to compare the very different force- extension curves for a single molecule in the two cases of a uniform (only f applied) and non-uniform (only g applied) stretch. In particular, taking advantage of our approximated formulas, we can analyse the case of a FJC and a WLC polymer. The 2D and 3D FJC results are plotted in Fig.6; on the other hand, the 2D and 3D WLC curves have been shown in Fig.7. For the WLC case we assumed κ = 10kBT for the bending modulus at T = 293K. This value is comparable to that of polymer chains of biological interest (e.g., for DNA κ = 15kBT ).12 In any case three curves have been reported for drawing all the possible comparisons: the response under the field g, the response under the force f = g and, finally, the re- sponse to an external force f = N g. Interesting enough we note that the curve corresponding to the field g is FIG. 7. (color online) Force-extension curves of a WLC poly- mer in an external field (or external force) with N=20. The red line corresponds to the approximated expressions given in Eqs.(46) and Eqs.(48) while the black circles have been ob- tained through MC simulations. The 2D (Eq.(44)) and 3D (Eq.(47)) WLC expressions (without an external field) are plotted for comparison with f = g and f = N g. The value of the bending spring constant is κ = 0.4 · 10−19 Nm ≃ 10kB T at T = 293K. always comprised between the cases with only the force f = g and f = N g. The response with the field g is clearly larger than that with the single force f = g since the field corresponds to a distribution of N forces (of in- tensity f ) applied to all monomers; therefore, the total force applied is larger, generating a more intense effect. However, the case with a single force f = N g shows a response larger than that of the field g. In this case the total force applied in the two cases is the same but the single force N f is applied entirely to the last terminal monomer, generating an overall stronger effect compared to the same force evenly distributed on the monomers. In fact, a force generates a stronger effect if it is placed in the region near the free polymer end (its effect is re- distributed also to all preceding bonds). The curves in Fig.6 and Fig.7 have been obtained with the theoretical formulations presented in this Section and confirmed by 9 TABLE I. Asymptotic forms of the force-extension curves for all cases described in the paper: FJC and WLC models in 2D and 3D geometry with force applied f or field applied g. Polymer chain Equation {z } FJC (2D) f Eq.(30) {z } FJC (3D) f Eq.(31) {z } FJC (2D) g Eq.(32) {z } FJC (3D) g Eq.(33) {z } WLC (2D) f Eq.(44) {z } WLC (3D) f Eq.(47) {z } WLC (2D) g Eq.(46) {z } WLC (3D) g Eq.(48) {z } Asymptotic form of r lN for f, g → 0 kB T(cid:17) (cid:16)x = lf kB T or lg 1 2 x 1 3 x Asymptotic form of r lN for f, g → ∞ kB T(cid:17) kB T or lg (cid:16)x = lf 1 − 1 2x 1 x 1 − 1 2 (cid:18)1 + N 2 (cid:19) x 1 3 (cid:18)1 + N 2 (cid:19) x log(N + 1) 2N log(N + 1) N 1 x 1 x 1 − 1 − Lp l x 2 3 Lp l x Lp l (cid:18)1 + N 2 (cid:19) x 1 − Lp l (cid:18)1 + 2 3 N 2 (cid:19) x 1 − 1 4 1 2 1 − 1 − 1 r Lp l x 1 r Lp l x 1 r Lp l x 1 l x r Lp √N + 1 − 1 2N √N + 1 − 1 N a series of MC simulations. In all case we obtained a quite perfect agreement between the two formulations. The knowledge of the closed-form expressions allowed us to analytically analyze the behavior of the chains for very low and very high applied forces (or fields). The results are shown in Table I: interestingly, we note that the ex- tension is always a linear function of the small applied perturbation. Nevertheless, the corresponding constant of proportionality depends on N only when a field is ap- plied to the chain; conversely, it is independent of N with a single force applied at one end. On the other hand, with a large perturbation applied to the molecule, we observe a 1/x behavior for the FJC models and a 1/√x behavior for the WLC models. To conclude we also remark that the order of the curves observed in Fig.6 and Fig.7 is con- firmed also in the low and high force (or field) regime by the following inequalities: 1 < 1 + N/2 < N (low force 10 regime) and 1 < log(N + 1) < N (high force regime) for the FJC model and 1 < 1 + N/2 < N (low force regime) and √N < 2(√N + 1 − 1) < N (high force regime) for the WLC model (always for N ≥ 2). V. ACTION OF A PULLING FORCE NOT ALIGNED WITH THE EXTERNAL FIELD In previous Sections we considered the polymer chain immersed in an external field with an external force equal to zero at its end. However, since we developed a form of the partition function also taking into account an exter- nal force applied at the end of the chain (at least for the FJC model), we can directly study the important case with a non zero force superimposed to an external field, in general having different orientation. To do this, we keep fixed the origin of the chain and apply a constant force at the end of the polymer with different angles with respect to the direction of the applied field. We will anal- yse such a problem for both the FJC and WLC cases. To begin, we consider a pulling force perpendicular to the direction of the applied field, respectively the y and z axis of our reference frame. For increasing values of the bending spring constant κ going from nearly zero (FJC model) to 8· 10−19Nm (WLC model, including the bend- ing constant of the DNA given by κ = 0.6 · 10−19 Nm ≃ 15kBT ). In Fig.8 we reported the results for the av- erage monomers positions and their variances. The red solid lines correspond to the analytical results for the FJC case, while the black symbols correspond to the MC simulations. It is interesting to observe the effect of the persistence length (or, equivalently of the bending stiff- ness): in fact, in the top panel of Fig.8 we note that the chains with an higher bending spring constant tend to remain more straight under the same applied load. At the same time, in the fourth panel of Fig.8 we ob- serve a decreasing variance along the z-axis (direction of the applied field) with an increasing bending spring con- stant; this fact can be easily interpreted observing that an higher rigidity of the chain reduces the statistical fluc- tuations in the direction of the applied field. The situ- ation is more complicated for the variances along the x and y directions: in fact, along the chain, there are some monomers with variances larger than the corresponding FJC case and others with smaller values. In Fig.9 the average positions of the monomers for dif- ferent directions of the external force are reported. The figure shows how the average monomer positions depend on the bending rigidity κ and on the external force angle θ. As before we can observe that the persistence length of the chain tends to maintain a low curvature in the shape of the chain. This phenomenon is more evident with an increasing angle between the force and the field. In fact, in Fig.9, the deviation between the FJC results and the WLC ones is higher for the angles approaching π, where the force and the field are applied in opposite directions. In Figs.10 and 11 the three components of the vari- i/l y , ri h 5 4 3 2 1 0 0 2.5 2 1.5 1 0.5 0 0 2.5 2 1.5 1 0.5 0 0 1 0.8 0.6 0.4 0.2 0 0 2 /l x , 2 i σ 2 /l y , 2 i σ 2 /l z , 2 i σ 5 10 15 hri,zi/l 0.5 i/N 1 0.5 i/N 1 0.5 i/N 1 FIG. 8. (color online) Action of a pulling force f (along the y-axis) perpendicular to the applied field g (along the z-axis). We adopted different values of the bending spring constant: κ = 0.08, 0.6, 2, 8 · 10−19 Nm. The chain length is fixed (N = 20), the external field amplitude is g = 4 pN and the force applied to the last monomer of the chain corresponds to f = 8 pN. The red solid lines correspond to the analytical results for the FJC case (see Eqs.(29) and (42)). Black circles correspond to the MC simulations with the different bending spring constants. In the top panel we reported the average positions, while in the others the three variances of the x, y and z components. 11 15 10 5 i/l y , ri h 0 −10 0 hri,zi/l 10 FIG. 9. (color online) Average positions of the chain for dif- ferent angles between the external traction force f and the direction of the applied field g. We adopted N = 20, g = 4 pN and f = 60 pN. The red solid lines correspond to the FJC ana- lytical result, Eq.(29). The symbols represent the MC results for the WLC model with κ = 0.08, 0.6, 2 · 10−19 Nm (circles, triangles and squares, respectively). For both FJC and WLC models we used different values of the angle between the ap- plied field and the traction force θ = π/2, 3π/4, 5π/6, 15π/16 from the right left. ance are reported versus the position of the monomer along the chain and the angle between the field and the force directions, for the FJC and WLC case, respectively. We can extract some general rules about this very com- plex scenario: as for the variance along the x direction we observe it to be an increasing function both of the po- sition i along the chain an of the angle θ between f and g. Both behaviors can be interpreted with the concept of persistence length, as discussed above. Conversely, the description of the variance along the y direction is more complicated. In fact, while the increasing trend of the variance with the position i along the chain is main- tained, we observe a non monotonic behavior in terms of the angle θ, with a minimum of the variance at about θ = 2π/3. Finally, the variance along the z direction is always increasing along the chain, but it shows a maxi- mum near θ = π (at least in the first part of the polymer chain). VI. CONCLUSIONS In this work we investigated mechanical and confor- mational properties of flexible and semi-flexible polymer chains in external fields. As for the FJC model we de- veloped a statistical theory, based on the exact analyt- ical determination of the partition function, which gen- eralizes previous results to the case where an external field is applied to the system. In particular we obtained closed form expression for both the average conformation of the chain and its covariance distribution. For sake of completeness, all calculations have been performed both in two-dimensional and three-dimensional geometry. On 0.5 0 −0.5 −1 −1.5 1 ) 2 /l x , 2 i σ ( g o l 0.5 0 −0.5 −1 −1.5 1 ) 2 /l y , 2 i σ ( g o l 0 ) 2 /l z , 2 i σ ( g o l −1 −2 −3 1 0.5 i/N 0 0 2 θ 0.5 i/N 0 0 2 θ 0.5 i/N 0 0 2 θ FIG. 10. (0 < θ < π) for the FJC model. As before we used N = 20, g = 4 pN and f = 60 pN. (color online) Monomer variances versus the position along the chain (i) and the angle between force and field 1 0 −1 1 ) 2 /l x , 2 i σ ( g o l 1 0 −1 1 ) 2 /l y , 2 i σ ( g o l 0 ) 2 /l z , 2 i σ ( g o l −1 −2 −3 1 0.5 i/N 0 0 2 θ 0.5 i/N 0 0 2 θ 0.5 i/N 0 0 2 θ FIG. 11. (color online) Monomer variances versus the position along the chain (i) and the angle between force and field (0 < θ < π) for the WLC model. As before we used N = 20, g = 4 pN and f = 60 pN. We also adopted a bending stiffness κ = 0.6 · 10−19 Nm. the other hand, as for the WLC model we derived new approximate expressions describing the force-extension curve under the effect of an external field. They can be considered as the extensions of the classical Marko-Siggia relationships describing the polymer pulled by a single external force applied at the free end of the chain. All our analytical results, for both FJC and WLC models, have been confirmed by a series of Monte Carlo simu- lations, always found in very good agreement with the theory. The overall effects generated on the tethered polymer by the application of an external field can be summarized as follows. As for the average configuration of a chain, it is well known that a single pulling force generates a uniform deformation along the chain (for a homogeneous polymer with all monomers described by the same effec- tive elastic stiffness). On the contrary, the application of an external field produces a non uniform deformation along the chain, showing a larger deformation in the por- tion of the chain closest to the fixed end. Moreover, the variances of the positions increase linearly along the chain with a single force applied to the polymer. Conversely, the polymer subjected to an external field exhibits a non- linearly increasing behavior of the variances along the chain. More specifically the variances assume the largest values nearby the last free monomers, where we can mea- sure the highest fluctuations. To conclude, we underline that the use of the MC method, once validated against known analytical solu- tions, is crucial for analysing models conditions which are beyond reach of a full analytical calculation. We take full profit of this approach for analysing the effects of the combination of an applied force at the free end together with an external field, especially when the two are not aligned. We have analysed the average configurational properties of the polymer, observing a very complex sce- nario concerning the behavior of the variances. ACKNOWLEDGMENTS We acknowledge computational support by CASPUR Italy) under project "Standard HPC Grant (Rome, 2011/2012". FM acknowledges the Department of Physics of the University of Cagliari for the extended visiting grant, and the IEMN for the kind hospitality of- 12 fered during part of this work. REFERENCES 1A. Ashkin, Proc. Natl Acad. Sci. 94, 4853 (1997). 2C. Gosse and V. Croquette, Biophys. J. 82, 3314 (2002). 3D. M. Czajkowsky and Z. Shao, FEBS Lett. 430, 51 (1998). 4C. R. Calladine, H. R. Drew, B. F. Luisi, and A. A. Travers, Understanding DNA: the molecule and how it works; Elsevier Academic Press, Amsterdam, 1992. 5A. Bensimon, A. Simon, A. Chiffaudel, V. Croquette, F. Heslot, and D. Bensimon, Science 265, 2096 (1994). 6E. Y. Chan, N. M. Goncalves, R. A. Haeusler, A. J. Hatch, J. W. Larson, A. M. Maletta, G. R. Yantz, E. D. Carstea, M. Fuchs, G. G. Wong, S. R. Gullans, and R. Gilmanshin, Genome Res. 14, 1137 (2004). 7M. Rief, F. Oesterhelt, B. Heymann, and H. E. Gaub, Science 275, 28 (1997). 8M. Rief, M. Gautel, F. Oesterhelt, J. M. Fernandez, and H. E. Gaub, Science 276, 1109 (1997). 9F. Manca, S. Giordano, P. L. Palla, F. Cleri, and L. Colombo, Two-state theory of single-molecule stretch- ing, submitted (2012). 10C. Bustamante, S. B. Smith , J. Liphardt, and D.Smith , Curr. Op. Struct. Biol. 10, 279 (2000). 11S. B. Smith, L. Finzi, and C. Bustamante, Science 258, 1122 (1992). 12J. F. Marko and E. D. Siggia, Macromolecules 28, 8759 (1995). 13F. Manca, S. Giordano, P. L. Palla, R. Zucca, F. Cleri, and L. Colombo, J. Chem. Phys. 136, 154906 (2012). 14J. M. Huguet, C. V. Bizarro, N. Forns, S. B. Smith, C. Bustamante, and F. Ritort, Proc. Nat. Ac. Sci. (PNAS) 107, 15341 (2010). 15D. W. Trahan and P. S. Doyle, Biomicrofluidics 3, 012803 (2009). 16S. G. Wang and Y. G. Zhu, Biomicrofluidics 6, 024116 (2012). 17C.-C. Hsieh and T.-H. Lin, Biomicrofluidics 5, 044106 (2011). 18D. C. Schwartz, Li X., L. I. Hernandez, S. P. Ram- narain, E. J. Huff, and Y.-K. Wang, Science 262, 110 (1993). 19T. R. Strick, J. F. Allemand, D. Bensimon, A. Bensi- mon and V. Croquette, Science 271, 1835 (1996). 20T. R. Strick, V. Croquette, and D. Bensimon, Nature 404, 901 (2000). 21C. Bustamante, J.C. Macosko, and G.J. Wuite, Nat. Rev. Mol. Cell. Biol. 1,130 (2000). 22D. E. Smith, H. P. Babcock, and S. Chu, Science 283, 1724 (1999). 23T. T. Perkins, D. E. Smith, R. G. Larson, and S. Chu, Science 268, 83 (1995). 24T. T. Perkins, S. R. Quake, D. E. Smith, and S. Chu, Science 264, 822 (1994). 25M. Warner, J. M. F. Gunn, and A. B. Baumgartner, J. Phys. A: Math. Gen. 19, 2215 (1986). 26G. J. Vroege and T. Odijk, Macromolecules 21, 2848 (1988). 27K. D. Kamien, P. L. Doussal, and D. R. Nelson, Phys. Rev. A 45, 8727 (1992). 28F. Brochard-Wyart, Europhys. Lett. 23, 105 (1993). 29F. Brochard-Wyart, H. Hervet, and P. Pincus, Euro- phys. Lett. 26, 511 (1994). 30F. Brochard-Wyart, Europhys. Lett. 30, 387 (1995). 31F. S. Henyey and Y. Rabin, J. Chem. Phys. 82, 4362 (1985). 32Y. Rabin, F. S. Henyey, and D. B. Creamer, J. Chem. Phys. 85, 4696 (1986). 33A. Lamura, T. W. Burkhardt, and G. Gompper, Phys. Rev. E 64, 061801 (2001). 34J. W. Gibbs, 1902. Elementary principles in statistical mechanics; Charles Scribner's Sons, New York, 1902. 35J. H. Weiner, Statistical mechanics of elasticity; Dover Publication Inc., New York, 2002. 36K. Binder, Rep. Progr. Phys. 60, 487 (1997). 37F. Manca, S. Giordano, P. L. Palla, F. Cleri and L. Colombo, J. Phys.: Conf. Ser. 383, 012016 (2012). 38A. E. Cohen, Phys. Rev. Lett. 91, 235506 (2003). 39J. R. Blundell and E. M. Terentjev, Soft Matter 7, 3967 (2011). 40L. Schwartz, Mathematics for Physical Sciences; Addison-Wesley, Reading, MA, 1966. 41M. Abramowitz and I. A. Stegun, Handbook of Mathe- matical Functions; Dover Publication Inc., New York, 1970. 42I. S. Gradshteyn and I. M. Ryzhik, Table of Integrals, Series and Products; Academic Press, San Diego, 1965. 43D. Frenkel and B. Smit, Understanding Molecular Sim- ulation; Academic Press, San Diego, 1996. 44M. P. Allen and D. J. Tildesley, Computer Simulations of Liquids; Clarendon Press, Oxford, 1987. 45J. Kierfeld, O. Niamploy, V. Sa-yakanit, and R. Lipowsky, Eur. Phys. J. E 14, 17 (2004). 46M. Rubinstein, R. H. Colby, Polymer Physics; Oxford University Press, New York, 2003. 47R. D. Kamien, Rev. Mod. Phys. 74, 953 (2002). 48N. J. Woo, E. S. G. Shaqfeh, B.Khomami, J. Rheol. 48, 281 (2004). 13
1011.1467
1
1011
2010-11-05T17:59:15
System Level Numerical Analysis of a Monte Carlo Simulation of the E. Coli Chemotaxis
[ "physics.bio-ph", "q-bio.CB" ]
Over the past few years it has been demonstrated that "coarse timesteppers" establish a link between traditional numerical analysis and microscopic/ stochastic simulation. The underlying assumption of the associated lift-run-restrict-estimate procedure is that macroscopic models exist and close in terms of a few governing moments of microscopically evolving distributions, but they are unavailable in closed form. This leads to a system identification based computational approach that sidesteps the necessity of deriving explicit closures. Two-level codes are constructed; the outer code performs macroscopic, continuum level numerical tasks, while the inner code estimates -through appropriately initialized bursts of microscopic simulation- the quantities required for continuum numerics. Such quantities include residuals, time derivatives, and the action of coarse slow Jacobians. We demonstrate how these coarse timesteppers can be applied to perform equation-free computations of a kinetic Monte Carlo simulation of E. coli chemotaxis. Coarse-grained contraction mappings, system level stability analysis as well as acceleration of the direct simulation, are enabled through this computational multiscale enabling technology.
physics.bio-ph
physics
SYSTEM LEVEL NUMERICAL ANALYSIS OF A MONTE CARLO SIMULATION OF THE E. COLI CHEMOTAXIS C. I. Siettos School of Applied Mathematics and Physical Sciences, National Technical University of Athens, GR 157 80, Greece ABSTRACT Over the past few years it has been demonstrated that “coarse timesteppers” establish a link between traditional numerical analysis and microscopic/ stochastic simulation. The underlying assumption of the associated “lift-run-restrict-estimate” procedure is that macroscopic models exist and close in terms of a few governing moments of microscopically evolving distributions, but they are unavailable in closed form. This leads to a system identification based computational approach that sidesteps the necessity of deriving explicit closures. Two-level codes are constructed; the “outer” code performs macroscopic, continuum level numerical tasks, while the “inner” code estimates –through appropriately initialized “bursts” of microscopic simulation- the quantities required for continuum numerics. Such quantities include residuals, time derivatives, and the action of coarse slow Jacobians. We demonstrate how these “coarse timesteppers” can be applied to perform equation-free computations of a kinetic Monte Carlo simulation of E. coli chemotaxis. Coarse-grained contraction mappings, system level stability analysis as well as acceleration of the direct simulation, are enabled through this computational multiscale “enabling technology”. Keywords: Bio-mechanics, Chemotaxis, Microscopic Models, Monte-Carlo simulations, Multi-scale Computations, Equation-Free approach 1. Introduction Chemotaxis, the process by which cells change their speed and/or orientate in space responding to environmental chemical gradients, has been thoroughly studied as it constitutes one of the basic survival and growth mechanisms of many micro-organisms. Revealing and understanding these mechanisms that pertain the function of this biological directional “kinesis” is of outmost importance in several fields of biology, biomechanics and medicine including homeostasis [Yi et al., 2000; Ito et al., 2004], ocean ecology[Stocker et al., 2008], biologically-inspired robotics [Long et al., 2004] and orthopedics [Lend et al., 1995; Poitout, 2004]. The chemotaxis pathway and biomechanics of E.coli have been well investigated through experiments [Macnab and Koshland, 1972, Brown and Berg,1974; Berg, 1975; Parkinson, 1993; Spiro et al., 2997; Waldo et al., 1999; Cluzel et al., 2000] providing useful insights also for other chemotactic species. A recent study provided that more than half of the bacteria have chemotactic ability [Wuichet and Zhulin, 2010]. The extensive study of E-coli has allowed the development of a many dynamical models ranging from the macro [Patlak, 1953; Keller and Segel, 1971a,b; Lapidus and Schiller, 1974; Biler, 1998; Dolak and Schmeiser, 2005 ] to the mechanistic-based micro-scale incorporating biological information even at the molecular inner-cell level [Alt, 1980; Rivero et al., 1989; Othmer and Stevens, 1997; Stevens, 2000; Hillen and Othmer, 2000; Painter et al., 2000; ]. For a detailed review of these models please refer to [Horstmann, 2003; Tindall et al., 2008; Hiller and Paiter, 2009]. No doupt, these mathematical models are invaluable tools that can help shedding light on still unknown issues regarding not only chemotaxis but also many other “adaptive” biochemical and genetic networks as well as human health [Wu, 2005; Wuichet and Zhulin, 2010] and life-cycle [Friedrich and Ju licher, 2007]. However, current modeling practice in the field is often characterized by the lack of accurate macroscopic deterministic models that quantitatively reflect the dynamic behavior of the systems modeled. Due to effectively stochastic nature and nonlinear complexity of the biological dynamics, state equations resulting at the macroscopic level from the appropriate balances are simply not available, or overwhelming difficult to derive. For example, under certain assumptions the dynamics at the macroscopic level can be represented by a Partial-Differential Equation (a Fokker-Planck) describing the evolution of the probability density, say, ( )xρ of the bacteria in space [Snitzer, 1993] : ( ) x ρ ∂ t ∂ ( ℘= x , , ρ ) p (1) p where ℘ is the partial differential operator and denotes the vector containing other “internal” variables. If cell devision or death is ruled out, the above equation reads [Othmer and Schaap, 1998]: ( ) xρ t ∂ ( )u D ρρ−∇⋅∇= (2) ∂ is the chemotactic velocity. u where D is the diffusion coefficient and u ρ . Over the years, various The above equation is not in a closed form as is not explicitly related to approximations have been proposed that express u in terms of other internal variables such as the gradient of the concentration of the attractant (or repellant). For example, a most common approximation, as extracted from kinetic theory is given by the Patlak-Keller/Segel relation reading [Patlak, 1953; Keller and Segel, 1971a,b]: c∇= χu (3) where χdenotes the chemotactic sensitivity, c is the concentration of the chemo-attractant (or repellant). The above closure is derived assuming that the cells emit immediately the chemical chemoattractants which are then diffused. Many other approaches for the calculation of the chemotactic sensitivity or the chemotactic velocity expressed in terms of other internal variables have been proposed (see the discussion in [Othmer and Schaap, 1998]. Yet, such closure relations are based on approximations and assumptions that introduce certain biases in the analysis. This prevents the performance of important computational tasks such as stability and bifurcation analysis, control and optimization, which rely on the availability of models, lumped or distributed, in terms of macroscopic (coarse) variables. On the other hand, good models of the underlying physics can be written at the microscopic/cellular level, and the modeling is performed via stochastic simulators (e.g. Monte Carlo, Brownian dynamics, Lattice- Boltzmann, Markov chains). However, at this fine biological level of representation, standard system analysis and design algorithms cannot be used efficient to handle the given detailed information. Towards this aim, it has been shown that “coarse timesteppers” establish a link between traditional numerical analysis and microscopic simulation [Gear et al., 2002; Makeev et al., 2002; Runborg et al., 2002; Kevrekidis et al., 2003; Siettos et al., 2003; Kevrekidis et al., 2004; Mooller et al., 2005; Moon et al., 2005; Russo et al., 2007]. The underlying assumption of this procedure (see also Figure 1) is that macroscopic models exist and can, in principle, be derived from microscopic ones (description at a much finer level) but they may lack a closed form description (e.g in terms of moments of the microscopic distributions) of their governing-coarse equations. This system identification based computational approach sidesteps the necessity of deriving good explicit closures, thus enhancing our efficiency in dealing with the problem in a systematic way. In [Setayeshgar et al., 2005] it is demonstrated how coarse-projective integration can be carried out using a microscopic-stochastic model of chemotaxis. In [Erban et al., 2006] coarse-projective integration is studied and analysed in detail for kinetic Monte-Carlo of random walks simulating bacterial chemotaxis. Here we demonstrate how “coarse timesteppers” can be applied to both perform (i) system level steady state and stability analysis and (ii) enable the acceleration of the evolution computations of a Markov chain-Monte Carlo simulation of the Esherichia coli (E.coli) chemotaxis. Through the proposed approach, we show, for the first time, that near equilibrium the coarse-grained dynamics of this system evolve on a two-dimensional slow manifold that can be parametrized by the mean and the variance of the microscopic distribution. The paper is organized as follows: in section 2 we present the Monte-Carlo model which describes the dynamics of the chemotaxis of E-Coli at micro level. In section 3 we present the proposed framework for multiscale computations. In section 4 we show how the concept of coarse timesteppers can be combined with system identification techniques in order to accelerate in time the detailed simulations. Section 5 provides the results of the analysis and we conclude in section 6. 2. Esherichia Coli Chemotaxis: A Biased Random Walk Model Bacteria, such as E. coli, have several flagella that help them direct their movements towards the most likely direction. Each flagellum rotates by the action of a rotary motor. A typical E-coli bacteria has up to 6 flagella. These can rotate either in a clockwise (CW) manner, causing the bacterium to tumble, or counter- clockwise (CCW), causing the bacterium to swim in a straight line. The overall movement of a bacterium is the result of altering tumble and swim phases. Tumbling frequency and duration time are most important factors of controlling the steering of the bacteria. When a decreasing concentration of an attractant or increasing concentration of a repellant is detected, the bacteria will increase the tumbling frequency and will thus increasing the probability of changing direction. A bacterium has three types of receptors (transmembrane proteins) for detecting changes in the concentration of attractants and repellents. The signals from these receptors are transmitted through the signal transduction system to the flagella motors, where Che proteins are activated. Increasing concentrations Che-Y protein, which is the output of this system, increases the tumbling frequency, and therefore bias the CW rotation. In general, the evolution of the changes in the transduction system can be represented by an evolution equation of the form: ,uFu d ( = dt nR∈u RS ∈ denotes the stimulus signal to the system. In most of the is the state vector and where systems, there is some kind of adaptation to the input signal: when S is time independent some functional of u should be also time-independent. In this paper a simple two dimensional ODEs model for excitation and adaptation is used [Spiro et al., 1997]: (1) )S (2a) ( ) ( ) tSf τ a − u 2 (2b) du 2 dt ( ) ( ) tSf + u 2 1 ) ( u − τ e du 1 dt = = ( ) 0 f f is the function encoding the signal transduction steps, and should satisfy the condition 0 = . The response of this simple model occurs in two time scales: the scale of excitation, which is characterized by eτ and the scale of adaptation, which is characterized by the time constant aτ . the time constant t >> αττ ,e , then whenever αττ <<e the model relaxes to If u = 2 ( ) Sf − u 1 2/N The probability that the cell is running (i.e the number of flagella rotating CCW is greater equal to , N being number of flagella) is given by the sum of the binomial probability mass function (voting hypothesis): (3) p CCW N = ∑ 2/ Nj = ⎛ ⎜⎜ ⎝ N j ⎞ ⎟⎟ ⎠ p j CCW ( 1 − p CCW jN − ) (4) 6=N , eq. (4) gives that the probability of CCW bias of a single bacteria is CCWp From experimental data we know that the probability bias, of a single flagella motor in the absence of stimulus, for a 2.95 μΜ concentration of Che-Y protein, is 0.64 [Spiro et al., 1997]. For The behavior of a single bacterium is determined by the following evolution rules [Spiro et al., 1997]: 1. Draw a random number ζ from a uniform random distribution on [ ]1 0 . 2. Compare ζ with the probability of switching direction of rotation: 880.= CCWp . p 1 Φ−= ± ~ k ± t Δ± (5) ±k tΔ where is the sampling time, are rational functions of the reaction rate constants and the CWP concentration. Che-Y. Cluzel, et al. (Science, 2000) have measured the equilibrium CW ( ) P and CCW bias ( CCW ) as a Hill function of the dissociation constant Kd and the concentration Y of Che-Y as follows: P CW = n Y n + n Y = k + + k + _k K d P CCW n Y n + n Y = k + + _k k + = K d (6a) (6b) . =n and 310. 13.=dK where 3a. IF a flagellum rotates CW AND ζ > p THEN keep rotating CW ELSE ENDIF Switch to CCW rotation 3b. IF a flagellum rotates CCW AND ζ > p THEN keep rotating CCW ELSE ENDIF Switch to CW rotation 4. IF the number of flagella that rotate CW is less than 3 THEN tumble ELSE ENDIF run IF previously running, THEN direction remains unchanged ELSE ENDIF direction = +/- 1, with equal probability 3. Multiscale computations of the E-Coli chemotaxis using coarse timestepping 3.1 Computational framework for steady-state and stability analysis We start with a brief overview of the “coarse timestepper”. In order to make clear the underlying concept, let us assume that due to the complexity of the system under study there are no explicit macroscopic equations in a closed form that can approximate the emergent macroscopic dynamics in an efficient manner. Yet, we assume that the physics are known in a more detailed level and thus we can develop a “good” microscopic computational model using simulation techniques such as Molecular dynamics, Monte-Carlo, Brownian dynamics and cellular automata. In general, the microscopic simulator can be represented by the following map: U S=+ 1 k T )pU ( , k , (7) uk puF ( ) , T=+ k 1 ( )k kT t k = tU U ≡ denotes the state vector of the microscopic distribution at time and k mR∈p N m N R R →× is the vector of the system’s parameters. is the time-evolution operator, where S :T R Hence, the microscopic simulator reports the values of the states of the microscopic distribution after an arbitrarily chosen macroscopic time interval T . Let us further assume that the macroscopic (system-level) dynamics of the system under study can be described by a map of the form n nR∈u :F R × R where denotes the macroscopic state vector, and is a smooth macroscopic time- T evolution operator. We furthermore assume that due to complexity of the problem under study such a macroscopic model is not available, or it is overwhelming difficult to derive in a closed form. When this is the case, a series of important system level tasks such as the tracing of coarse-grained solution branches in the parameter space, stability analysis, optimization and design of control systems cannot be performed by exploiting the arsenal of well-established techniques in the continuum. The question that naturally arises is how one can systematically study and analyze the macroscopic (system-level) dynamics, when the macroscopic evolution operator is not explicitly available in a closed form. The answer comes from the concept of timestepping which constitutes the “heart” of the Equation- Free computational approach [Kevrekidis et al., 2003, Comm. Math. Science, 1,715; Makeev et al., 2002, J. Chem. Phys., 116, 10083 ; Gear et al., 2002, Comp. Chem. Engng. 26, 941 ; Siettos et al., 2003, J. Chem. Phys., 118, 10149]. The main idea is to sidestep the derivation of system-level equations and to use appropriately-initialized short runs in time of the microscopic/stochastic models to estimate necessary quantities “on demand”. What the Equation-Free approach does, in fact, is providing closures “on (8) , p demand”; relatively short bursts of the fine scale simulator naturally establish in a strictly numerical manner the slaving relation between the fast and slow dynamics of the system under study [refer to Gear et al., 2002, Comp. Chem. Engng. 26, 941; Siettos et al., 2003, J. Chem. Phys., 118, 10149; Kevrekidis et al., 2003, Comm. Math. Science, 1, 715 for more detailed discussions]. In a nutchel, the coarse timestepper consists of the following steps (see also figure 1): ( )0tu (a) Prescribe a macroscopic initial condition (e.g. concentration profile) (b) Transform it through lifting to one (or more) fine, consistent microscopic realizations ( ( )0 ) , where μ is the lifting operator. u U t t μ= 0 (c) Evolve this(ese) realization(s) using the microscopic simulator for the desired short ( )TU macroscopic time (time horizon) T , generating the value(s) . An appropriate choice T can be estimated from the spectrum derived of the mean field equations linearization around their steady states. ( )TMT ( ) u U (d) Obtain the restrictions = Ι=Μμ operators should satisfy . ; M denotes the restrict operator. The lift and restrict ; The above procedure can be considered as a “black box” coarse timestepper u =+ 1 k uΦ ( T k p ) , (9) Under certain assumptions regarding the separation between the time-scales of the system [Kevrekidis et al. TΦ n p n TF R xRR → 2003], . provides an numerical approximation of the unavailable : Thus: (e) At an outer level, and depending on the task we want to carry out, (such as the computation of fixed points, the stability analysis and optimisation), well established numerical analysis algorithms can be utilized to estimate “on demand” the required quantities such as residuals, Jacobians, control matrices and Hessians. These algorithms call the timestepper as a black-box subroutine from nearby appropriately perturbed initial conditions and for relatively short time intervals. For example, coarse-grained (macroscopic) steady states can be obtained as fixed points, using T as sampling time, of the mapping ΤΦ : u,Φu ( − T p ) = 0 (10) using Newton-Raphson method [Kelley, 1995]. The procedure involves the iterative solution of the following linearized system: )] (11) [ ( puΦu , −−= Τ − I ⎡ ⎢ ⎣ Φ ∂ Τ u ∂ ⎤ u δ ⎥ ⎦ The computation of the Jacobian Φ ∂ Τ can be achieved in a fully numerically manner by calling the u ∂ coarse timestepper from appropriately perturbed initial conditions. For example, using center finite ( ) j i , differences the element of the Jacobian is given by: Φ ∂ jΤ u ∂ i = ( uΦ jΤ + i ( u ε − i ( u ε i ) ) ) ( ) uΦ − i jΤ ( ) u 2 ε i () ( uO ε i 2 ) + (12) u th th i − j − and ΤΦ and elements of respectively; ε is a small and iu jΤΦ where denote the , appropriately chosen scalar. The above framework enables the temporal simulator to converge to both stable and unstable solutions and trace their locations through bifurcation points utilizing techniques such as the pseudo-arc length continuation approach. For large-scale systems, fixed point calculations can be performed in a more efficient manner using matrix- free iterative solvers such as the Newton-Generalized Minimum Residual (Newton-GMRES) method [Kelley, 1995]. The advantage of using matrix-free methods over more “traditional” techniques is that the Φ ∂ Τ is not required. Instead, what is really needed is u ∂ matrix-vector multiplications which can be obtained at low computational cost by calling the timestepper TΦ from nearby initial conditions allowing the estimation of the action of the linearization of the map on known vectors, as explicit calculation and storage of the Jacobian uΦ ( T + uΦv )( -) ε T ε vuΦ TD )( ≈⋅ (13) Alternative algorithms such as the Recursive Projection method [Shroff & Keller, 1993] and other Newton- Picard methods such as the ones presented in [Lust et al., 1998; Kavousanakis et al., 2008] can be also used to compute both steady states and periodic solutions and construct their bifurcation diagrams. TD Φ The leading (algebraically largest) eigenvalues of the matrix determine the local stability of the system. For large-scale systems these can be estimated using again a matrix-free iterative eigensolver such as the Arnoldi procedure [Saad, 1992; Christodoulou & Scriven 1998] exploiting the same timestepper approach. These algorithms share common procedures that in short can be described as follows (see also figure 2): using (9) j ,v vΦ t D , ⋅ =〉 T t j , 21 ,...,j nR∈1v 1 =v Choose with For j =1 Until Convergence (1) Compute and store 1 , j (3) vF ⋅ TD h 〈= (2) Compute and store t j ∑ t 1 = / 21 〉 〈= / r vΦ ⋅ T r j h , j j 1 + + = 1 h v r , h t D = − v r j j 1 + , j , j t j j j j (4) (5) End At step j ,the algorithm produces an orthonormal basis { v,v 1 2 ,...., }mv of the Krylov subspace jK spanned by { vΦ,v D ⋅ T 1 1 ,...., D Φ The projection of TD F on jK is represented in the basis }1 − j j v 1 − ⋅ T { }jv by the upper Hessenberg matrix (14) H j = v T j D vΦ ⋅ T j whose elements are the coefficients ijh . At each j step, GMRES minimizes the residual R T u,Φu−= ( p ) . (15) Regarding the stability analysis, the eigenvalues of jH provide approximations of the TD Φ for the outermost spectrum of TD Φ , where the eigenvectors of D TΦ are approximated by x = j zv j , j (16) Where jz are the eigenvectors of the Hessenberg (it may be computed using standard packages like EISPACK or LAPACK). j (uΦ T + (Φ)v − ε j T ε u ) The approximation character of the algorithm, flashes a note of caution in the choice of the perturbation parameter ε since at the end of the algorithm what is usual done is the formation of the converged Hessenberg through the multiplication T j v = A H )v j (17) T (v ≈ j which is only an approximation to the actual Hessenberg. This may reflect to some “distortion” to the computation of the leading eigenvectors. One should chooseε according to the accuracy of the already converged solution. 3.2 The lift and restrict operators for the chemotaxis problem As explained in 3.1, we don’t try to find any closures. Instead we exploit the concept of coarse timestepping to sidestep the derivation of such approximations. For our analysis, the spatial distribution in ( )xρ x , is computed by approximating the corresponding inverse cumulative distribution function ( ρ ICDF ) sn q (see also figure 2). For orthogonal polynomials or splines of order using for example )snm × ension ( W the matrix of dim ue by containing the val s of such example let us denote ( )ρ sn orthogonal polynomials over the m points in space ICDF h the on whic is computed [Setayeshgar t al., 2005]. e The restrict operator is defined as the product of a orthogonal polynomials i,.e.: since the lift operator creates WW =T ICDF a I ( )ρ ICDF ( )ρ with W resulting to the coefficients of the = ICDF ( ) Wρ ⋅ (i.e. a distribution in space) as: ICDF ( ) ρ TWa ⋅= (18) (19) 4. Coarse Projective Integration of the Markov Chain Monte Carlo Chemotaxis Model. The coarse timestepper can be also used to perform coarse projective integration (see Fig. 3). The basic idea is that coarse time-stepper can be used to approximate the time derivatives of the corresponding continuum formulation, even if the continuum equations are not known in closed form. Specifically, we execute the following steps: ( )itU , as well as their restrictions (f) Repeat step (d) over several time steps, giving several ( ) ( ) u U t i tM k , ...., ,2 ,1 1 . = = + i i (g) Use the chord connecting these successive time-stepper output points to estimate the derivative of the continuum variables. Note that this does not require that we know the explicit continuum equations. (h) Use this derivative in an outer integrator (such as forward Euler) to estimate the continuum ( ) u kt state much later in time. m ++1 (i) Go back to step (b). For the chemotaxis problem, the proposed procedure can be summarized in a two-tier level pseudo-code as follows: Do while {desired time Tfinal} I. For t = n, n+1, n+2, …n+l (a) Compute the Cumulative distribution function of x-positions ( )t PDF ρ from the probability distribution function Approximate the Inverse Cumulative Distribution function of x-positions ( )t ICDF ρ using sn orthogonal polynomials. Figure 4 shows five such orthogonal polynomials. Restore the coefficients polynomials. of the approximating ( )t CDF ρ [ 1=a aa , α ,..., ns (b) (c) ]t 2 End For II. Project the polynomial coefficients k-times instances ahead to estimate the ( ) ++ρ ICDF . kl t Several techniques can be used for that purpose, including ARMAX models [Ljung, 1987] and polynomial curve fitting using Least Squares; At this point we can also incorporate discrete time filters to smooth the time-series signal. This step involves the selection of an appropriate model structure (among a set of candidate models). Here, we demonstrate an Autoregressive (AR)-based system identification technique: (a) (b) (c) + c i = s − 2 n k + − = kn − ... n ..., ,2 ,1 the approximating polynomial identification of Perform system coefficients of the ICDF by a discrete model of the form ) ) ( ( ( ( ) ) ta tac tac ta 1 , +− i i i i 2 1 ia The above expression relates at time t to a finite number of past values ( ) nta . At this point we should note that, if necessary, a smoothing − i k filter can be applied to the a time series. i In this work, we implemented a second order (in a window of 5 sample points), Savitzky Golay filter [Savitzky and Golay, 1964], which is a time varying FIR filter. Savitzky-Golay smoothing filters (also called digital smoothing polynomial filters or least squares smoothing filters) are typically used to “smooth out” a noisy signal whose frequency span (without noise) is large. Savitzky-Golay filters are optimal in the sense that they minimize the least-squares error in fitting a polynomial to each noisy frame of data. The estimation of the vector of [ ] T =Θ c1 , .. ., c at time t is carried out using least parameters n-k squares (LS) over the “raw” or smoothed data at time instances t = n, n+1, …n+l. Use the derived model to project ai to time t+l+k. Use the projected values to lift to one (or more) consistent microscopic x distributions of positions End While e S = chemoatractant profile is a Gaussian-like function given by It is important to note that one must integrate the microscopic rules for some time before estimating the time derivative of the continuum variables. This allows higher moments of the continuum description to become slaved to the statistics of interest. 5. Numerical Analysis Simulation results were obtained using 2000 cells (Ncells) and dt = 0.1 as the Monte Carlo time step. The 2 x ( 6) − 1 − 2 π Figure 5b illustrates the left and right moving distributions for the Gaussian attractant profile. The distributions were derived by averaging over time from t=10000 s to t=15000 time steps using a time horizon of T=100 time steps (i.e averaging over a total of 50 samples) and over the total distribution of positions. The initial condition was a uniform (flat) profile of positions. As it can clearly shown, simulation results are consistent with those dictated by the voting hypothesis giving a probability of ~ 0.88 that the cell is running either left or right. Figure 6 depicts the distributions of the number of left moving flagella that rotate CCW (figure 6a) and tumbling cells (Figure 6b). Simulation results are consistent with those dictated by the binomial probability mass functions. For the problem under study the analysis is based on the hypothesis that a macroscopic coarse model exists ( )xρ and closes for the microscopic distribution of bacteria positions . This implies that all the other “inner” variables, such as number of flagellea rotating CCW or CW, the excitation u1 and adaptation u2 signals, become quickly slaved to the spatial distribution (they evolve towards a “slow manifold” parameterized by the underlying microscopic distribution). Computational results corroborating this can be seen in Fig 7, which illustrates the relatively fast slaving of the “internal” variables. Here we initialized the bacteria by setting u1 = u2 = 0 for all cells, running right with 4 flagellae rotating CCW. (Figure 5a). A = . The eigenvalues of the 2x2 matrix convergence, the Jacobian was found to be ICDF For the coarse-projective integration, the approximation of is made with linear extrapolation in time, l = 10 and k = 10 using eight orthogonal basis functions until t = 6000 time steps and then l =10, k = 20 till t = 25000 time steps, while after each lifting a period of 5T is used as “healing” of the errors made in the lifting step before the acquisition of the training data. In figure 8, are given the evolution of both normal integration (blue lines) and coarse-projective integration (red lines) for the density and cumulative distribution functions. The steady-state and stability analysis was performed using two different methods, namely (i) through Newton-Raphson’s method and (ii) through matrix-free methods. Newton-Raphson’s method was applied in terms of the mean value and the variance of a normal distribution function. Under this assumption, lifting was done with the inverse normal distribution. The fixed point was evaluated through the coarse timestepper with T= 400 time steps, 10 copies of cell distributions. Upon convergence of the Newton-Raphson to a residual of O(10-3), for perturbations ε ~5x10-2) the mean value was found to be ~6.005 and the final variance ~ 0.13. The values of the mean and standard deviation of the normal distribution in each step of Newton’s iteration is shown in figure 9b. The initial values of the mean and standard deviation of the distribution at t = 10000 time steps where we started Newton are also shown. Newton-Raphson’s method converged in 3 iterations (see figure 10a). Upon 0.90521 0.0010 - ⎤ ⎡ ⎢ ⎥ 0.012 - 0.8423 ⎣ ⎦ are: λ1~0.91 and λ2~0.845. Hence the corresponding eigenvalues in the s-plane are: λ1 =-0.023 and λ2 =- 0.0421. Looking at the Jacobian, the interaction of the non-diagonal terms is very small when compared with the diagonal terms. This is an indication that near the fixed point the particular system can be considered for practical purposes uncoupled. Newton-GMRES was used to compute the steady states with respect to the coefficients of the orthogonal basis functions and Arnoldi’s method has been employed to perform stability analysis. These matrix-free algorithms were wrapped around the microscopic simulator with T=2000 time steps, averaging over 20 copies of cells distributions. For the approximation of the inverse cumulative distribution function we used eight orthogonal basis functions and we asked for the 2-leading eigenpairs with a residual of O(10-4). For comparison purposes the eigenvalues (in the s-plane) as calculated using Newton-Raphson’s method with lifting through the inverse normal distribution function and the Arnoldi’s method using eight orthogonal basis functions are given in table 1. Figure 10b shows the eigenvectors corresponding to the symmetric (variance of distribution of bacteria positions) and antisymmetric (mean of the distribution of bacteria positions) part of perturbation. As it is shown, the leading eigenvalues and their corresponding eigenvectors as computed with Newton- Raphson and Arnoldi coincide for any practical means. This reveals that near the equilibrium the coarse- grained dynamics evolve on a two dimensional manifold that can be in principle parameterized by the first two moments of the underlying microscopic distribution, namely the mean and the variance of the spatial distribution. 6. Conclusions The Equation-free approach is a computational framework that provides a systematic approach for analyzing the parametric behavior of complex/ multiscale simulators much more efficiently than simply simulating forward in time. Acceleration of simulations in time, regime and stability computations, as well as continuation and numerical bifurcation analysis and other important tasks such as rare-events analysis of the complex-emergent dynamics can be performed in a straightforward manner bypassing the explicit extraction of closures. In this work, we have demonstrated how this multiscale approach can be used to extract system-level information from a Monte-Carlo model simulating the chemotaxis of E. COLI. Fixed point and stability analysis were performed using both “conventional” contraction maps such as Newton-Raphson and matrix- free iterative algorithms (Newton-GMRES for fixed point computations and Arnoldi eigensolver). Through the numerical analysis we showed that near the equilibrium the coarse-grained dynamics evolve on a two dimensional manifold that can be defined as a function of the first two moments of the evolving spatial distribution. Coarse projective integration was also demonstrated combining the concept of coarse- timestepping with nonlinear system identification techniques. The particular one-dimensional in space stochastic model does not exhibit any striking nonlinear behavior such as the appearance of critical points, phase transitions or blow-up solutions in finite time. Such phenomena including complex pattern formation have been observed in experimental-observed chemotactic responses [Adler and Templeton, 1967; Budrene and Berg, 1991, 1995] and their dynamics have been approximated and analysed using macroscopic level models [Keller and Segel, 1971a,b; Brenner et al., 1998; Myerscough et al., 1998; Polezhaev et al., 2006] . To this end we believe that the proposed framework can be used to deepen our understanding in the way chemotaxis causes the emergent of such complex behaviour. Systematic coarse- grained bifurcation and stability analysis for both steady state and periodically oscillatory solutions [Russo et al., 2007; Kavousanakis et. al., 2008], the adaptive detection of the critical points that mark the onset of such phenomena [Siettos et al., 2006] can be efficiently performed exploiting the detailed biological knowledge that is incorporated into state-of-the art-microscopic models. Acknowledgments The author would like to thank Professor Yannis Kevrekidis (Dept. of Chemical Engineering, Dept. of Mathematics and PACM, Princeton University), Professor C. W. Gear (Dept. of Chemical Engineering, Princeton University) and Professor Hans Othmer (Dept. of Mathematics, University of Minnesota) for their collaboration, insightful comments and many fruitful and helpful discussions. References Adler, J., Templeton, B., The effect of environmental conditionson the motility of Escherichia coli. J. Gen. Microbiol. 46, 175–184 (1967). Alt, W., Biased random walk model for chemotaxis and related diffusion approximation, J. Math. Biol. 9, 147–177 (1980). Berg H.C, Chemotaxis in bacteria, Annu Rev Biophys Bioeng. 4, 119-136 (1975). Biler, P., Local and global solvability of some parabolic systems modelling chemotaxis, Adv. Math. Sci. Appl. 8, 715–743 (1998). Brenner, M. P., Levitov, L. S., Budrene, E. O., Physical Mechanisms for Chemotactic Pattern Formation by Bacteria, Biophysical J. 74, 1677-1693 (1998). Brown, D., Berg, H., Temporal stimulation of chemotaxis in Escherichia coli. PNAS 71, 1388–1392 (1974). Budrene, E., Berg, H., Complex patterns formed by motile cells of Escherichia coli, Nature 349, 630–633 (1991). Budrene, E., Berg, H., Dynamics of formation of symmetrical patterns by chemotactic bacteria, Nature. 376, 49-53 (1995). Christodoulou, K. N., Scriven, L. E., Finding leading modes of a viscous free surface low: an asymmetric generalized problem, J. Sci. Comp. 3, 355-406 (1998). Cluzel, P., Surette, M., Leibler, S., An Ultrasensitive bacterial motor revealed by monitoring signaling proteins in single cells, Science 287, 1652-1655 (2000). Dolak, Y., Schmeiser, C., The Keller-Segel model with logistic sensitivity function and small diffusivity, SIAM J. Appl. Math. 66, 286–308 (2005). Erban, R. Kevrekidis, I. G., Othmer, H. G., An equation-free computational approach for extracting population-level behavior from individual-based models of biological dispersal, Physica D 215, 1-24 (2006). Friedrich, B. M., Jülicher, F., Chemotaxis of sperm cells, PNAS 104, 13256-13261 (2007). Gear, C. W., Kevrekidis, I. G. & Theodoropoulos, C., Coarse Integration/Bifurcation Analysis via Microscopic Simulators: micro-Galerkin methods, Comp. Chem. Eng. 26, 941-963 (2002). Hillen, T., Othmer, H.G., The diffusion limit of transport equations derived from velocity jump processes, SIAM J. Appl. Math. 61, 751–775 (2000). Hillen, T., Painter, K. J., A user’s guide to PDE models for chemotaxis, J. Math. Biol. 58, 183–217 (2009). Horstmann, D., From 1970 until present: the Keller–Segel model in chemotaxis and its consequences I, Jahresberichte DMV 105, 103–165 (2003). Ito, M., Xu, H., Guffanti , A. A., Wei, Y., Zvi, L., Clapham, D. E., Krulwich, T. A., The voltage-gated Na+ channel NaVBP has a role in motility, chemotaxis, and pH homeostasis of an alkaliphilic Bacillus, PNAS 101, 10566-10571 (2004). Kavousanakis, M., Russo, L., Siettos, C. I., Boudouvis, A. G., Georgiou, G. C., A Timestepper Approach for the Systematic Bifurcation and Stability Analysis of Polymer Extrusion Dynamics, J. Non-Newtonian Fluid Mechanics 151, 59-68 (2008). Keller, E., Segel, L, Initiation of slime mold aggregation viewed as an instability, J. Theor. Biol. 26, 399– 415 (1970). Keller, E.F., Segel, L.A.: Model for chemotaxis. J. Theor. Biol. 30, 225–234 (1971a). Keller, E.F., Segel, L.A., Traveling bands of chemotactic bacteria: a theoretical analysis, J. Theor. Biol. 30, 377–380 (1971b). Kelley, C. T., Iterative Methods for Linear and Nonlinear Equations, SIAM, Philadelphia, 1995. Kevrekidis, I. G., Gear, C. W., Hummer, G., Equation-free: the computer-assisted analysis of complex, multiscale systems, AI.Ch.E. J. 50, 1346-1354 (2004). Kevrekidis, I. G., Gear, C. W., Hyman, J. M., Kevrekidis, P. G., Runborg, O. & Theodoropoulos, C., Equation-free coarse-grained multiscale computation: enabling microscopic simulators to perform system- level tasks, Comm. Math. Sciences 1, 715-762 (2003). Lapidus, R., Schiller, R., A mathematical model for bacterial chemotaxis, Biophys. J. 14, 825–834 (1974). Lind, M., Deleuran, B., Thestrup-Pedersen, K., Soballe, K., Chemotaxis of human osteoblasts: effects of osteotropic growth factors, APMIS 103, 140-146 (1995). Ljung, L., System identification: theory for the user, Prentice Hall, Englewood Cliffs, 1987. Long, J.H., Lammert, A.C., Pell, C.A., Kemp, M., Strother, J.A., Crenshaw, H.C., McHenry, M.J., A navigational primitive: biorobotic implementation of cycloptic helical klinotaxis in planar motion , IEEE J Oceanic Eng. 29, 795–806 (2004). Lust, K., Roose, D., Spence, A., Champneys, A. R., An adaptive Newton-Picard algorithm with subspace iteration for computing periodic solutions, SIAM J. Sci. Comp 19, 1188-1209 (1998). Macnab, R. M., Koshland, D.E., The Gradient –Sensing Mechanism in Bacterial Chemotaxis, PNAS 69, 2509-2512 (1972). Makeev, A., Maroudas, D. & Kevrekidis, I. G., Coarse stability and bifurcation analysis using stochastic simulators: Kinetic Monte Carlo Examples, J. Chemical Physics. 116, 10083-10091 (2002). Mooller, J., Runborg, O., Kevrekidis, P. G., Lust, K., Kevrekidis, I. G., Equation-free, effective computation for discrete systems: A time stepper based approach, Int. J. of Bifurcation and Chaos 15, 975- 996 (2005). Moon, S. J., Ghanem, R., Kevrekidis, I. G., Coarse graining the dynamics of coupled oscillators, Phys. Rev. Let. 96, 144101, 1-4 (2006). Myerscough, M.R., Maini, P.K., Painter, K.J., Pattern formation in a generalized chemotactic model. Bulletin of Mathematical Biology 60, 1—26 (1998). Othmer, H.G., Schaap, P., Oscillatory cAMP signaling in the development of Dictyostelium discoideum, Comments on Theoretical Biology 5, 175-282, (1998). Othmer, H.G., Stevens, A., Aggregation, blowup and collapse: The ABC’s of taxis in reinforced random walks, SIAM J. Appl. Math. 57, 1044–1081 (1997). Painter, K.J., Maini, P.K., Othmer, H.G.: Complex spatial patterns in a hybrid chemotaxis reaction- diffusion model, J. Math. Biol. 41, 285–314 (2000). Parkinson, J.S., Signal transduction schemes of bacteria, Cell 73, 857-871 (1993). Patlak, C.S., Random walk with persistence and external bias, Bull. Math. Biophys. 15, 311–338 (1953). Poitout, D. G., Biomechanics and Biomaterials in orthopedics, Springer, 2004. Polezhaev, A. A., Pashkov, R. A., Lobanov, A. I., Petrov, I. B., Spatial patterns formed by chemotactic bacteria Escherichia coli, Int. J. Dev. Biol. 50, 309-314 (2006). Rivero, M., Tranquillo, R., Buettner, H., Lauffenburger, D., Transport models for chemotactic cell populations based on individual cell behavior, Chem. Eng. Sci. 44, 2881–2897 (1989). Runborg, O., Theodoropoulos, C. & Kevrekidis, I. G., Effective Bifurcation Analysis: a Time-Stepper- Based Approach, Nonlinearity 15, 491-511 (2002). Russo, L., Siettos, C. I., Kevrekidis, I. G., Reduced Computations for Nematic-Liquid Crystals: A Timestepper Approach for Systems with Continuous Symmetries, J. Non-Newtonian Fluid Mechanics 146, 51-58 (2007). Saad, Y., Numerical methods for large eigenvalue problems, Manchester University Press: Oxford- Manchester, 1992. Savitzky, A., Golay, M. J. E., Smoothing and differentiation of data by simplified least-squares procedures, Anal. Chem. 36, 1627-1639 (1964). Schnitzer, M. J., Theory of continuum random walks and application to chemotaxis, Physical Review E 48, 2553-2568 (1993). Setayeshgar, S., Gear, C. W., Othmer, H. G., Kevrekidis, I. G., Application of coarse integration to bacterial chemotaxis, Multiscale Mod. Sim. 4, 207-327 (2005). Shroff, G. M. & Keller, H. B., Stabilization of Unstable Procedures - the Recursive Projection Method, SIAM J. Num. Anal. 30, 1099-1120 (1993). Siettos, C. I., Graham, M. & Kevrekidis, I. G., Coarse Brownian Dynamics for Nematic Liquid Crystals: Bifurcation Diagrams via Stochastic Simulation, J. Chemical Physics 118, 10149-10156 (2003). Siettos, C.I., Rico-Martinez, R., Kevrekidis, I. G., A Systems-Based Approach to Multiscale Computation: Equation-Free Detection of Coarse-Grained Bifurcations, Computers Chem. Eng. 30, 1632-1642 (2006). Spiro, P. A., Parkinson, J. S, Othmer, H. G., A model of excitation and adaptation in bacterial chemotaxis, PNAS 94, 7263–7268 (1997). Stevens, A., The derivation of chemotaxis-equations as limit dynamics of moderately interacting stochastic many particle systems, SIAM J. Appl. Math. 61, 183–212 (2000). Stocker, R., Seymour, J. R., Samadani, A., Hunt, D. E., Polz, M. F., Rapid chemotactic response enables marine bacteria to expoit ephemeral microscale nutrient patches, PNAS 105, 4209-4214 (2008). Waldo, G. S., Standish, B. M., Berendzen, J., Terwilliger, T. C., Rapid protein-folding assay using green fluorescent protein, Nature Biotechnol. 17, 691-695 (1999). Wu, D., Signaling mechanisms for regulation of chemotaxis, Cell Res. 15, 52–56 (2005). Wuichet, K. and Zhulin, I. B., Origins and diversification of a complex signal transduction system in prokaryotes, Sci. Signal., 3, ra50 (2010) Yi, T-M., Huang, Y., Simon, M. I., Doyle, J., Robust Perfect adaptation in acterial chemotaxis through integral feedback control, PNAS 97, 4649–4653 (2000). Figures Microscopic Microscopic Simulator Simulator Fixed Po int Solver Fixed Po int Solver Fixed Po int Solver Iterat ive Iterat ive Iterat ive Iterat ive EigenSolver EigenSolver EigenSolver EigenSolver Res tric t Res tric t M M Lift Lift μ μ Statio nary solutions Statio nary solutions Statio nary solutions Statio nary solutions Stabilit y Anal ysi s Stabilit y Anal ysi s Stabilit y Anal ysi s Stabilit y Anal ysi s Chemoattractant Chemoattractant Chemoattractant Profile Profile Profile Figure 1: The coarse timestepper x = x(CDF) x = x(CDF) Microscopic Simulator Coarse timestepper 0 [ a 1 a 2 0 1 Lift μ [ a 1 a 2 ... nsa ] (t) Figure 2. Lifting and Restriction of the microscopic distribution 1 Restriction M ] (t nsa ... + 1) (CDF) x = x x = x (CDF) x = x (CDF) x = x (CDF) 1 0 ... nsa += kt 2 Microscopic Simulator 0 1 nsa kt 2 ++= M a 1 a 2 ... Microscopic Simulator 1 nsa Microscopic Simulator 0 1 ... += kt 1 a 1 a 2 0 a 1 a 1 a 2 a 2 nsa = kt ... Figure 3: Coarse projection 3 2 1 0 -1 -2 -3 0 500 1000 1500 2000 Figure 4. Orthogonal basis functions used to approxim ate the ICDF. 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 2 4 6 8 10 12 0.5 0.4 0.3 0.2 0.1 0 5.6 5.8 6 6.2 6.4 (a) Figure 5. (a) Gaussian attractant profile, (b) Distributions of bacteria: Running left < , right > and tumble o. (b) 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 8 10 2 4 6 8 10 6 (a) 0.4 0.3 0.2 0.1 0 2 4 Figure 6. (a) Flagella rotating CCW Distributions for the left moving cells; triangles-up correspond to 4, asterics to 3, circles to 5 and triangles-down to 6 CCW rotating flagella (b) Flagella rotating CCW Distributions for the tumbling cells-Flat attractant distribution. Squares correspond to 2 and rhombs to 1 CW rotating flagella. 1 0.8 0.6 0.4 0.2 0 0 10 20 (x10) s 30 40 1u 4 3 2 1 0 -1 15 10 time 5 0 7 5.5 6 x 6.5 (a) (b) Figure 7. Quick dynamic slaving of “internal” variables. When lifting, (i) we set u(1) = u(2)= 0 for everybody and (ii) we set all cells running with 4 flagellae CCW (iii) we start with all cells running RIGHT. (a) Time evolution of number of cells running right (circles), running cells (either left or right) having 4 flagella rotating CCW (triangles right), tumbling cells (asterisks), (b) evolution of variable . (Gaussian attractant profile) 1u (a) (b) Figure 8. Coarse projection (5 healing, m=5 acquisition, k=10 projection & 8 basis functions till time=6000 and then 5 healing, 5 acquisition, k=30 projection & 8 basis functions till time=25000). Blue lines correspond to temporal simulations while red ones to coarse projected simulation. 6.2 6.15 6.1 6.05 6 5.95 5.9 1 0.24 0.22 0.2 0.18 0.16 0.14 0.12 1 2 3 4 5 2 3 4 5 igure 9. (a) Mean and (b) standard deviation of the normal distribution in each step of Newton’s F iteration. 0.025 0.02 0.015 0.01 0.005 1 2 3 4 5 0.3 0.2 0.1 0 -0.1 -0.2 -0.3 0 500 1000 1500 2000 igure 10. (a) Newton’s convergence (error-iterations) (b) System Eigenvectors of the ICDF; F solid lines correspond to eigenvectors as computed with Newton’s method while the dotted ones correspond to the eigenvalues as computed with Arnoldi’s methos; from left: upper (bottom) lines correspond to λ1~0.61 (λ2~0.44). ARNOLDI λ1=-0.025 λ2=-0.041 λ1=-0.023 λ2=-0.042 N EWTON-RAPHSON able. 1. Eigenvalues as calculated using Newton’s method with lifting through the inverse normal T distribution function and the Arnoldi’s method using 5 orthogonal basis functions.
1503.08150
4
1503
2015-08-20T13:05:32
Probing protein orientation near charged nanosurfaces for simulation-assisted biosensor design
[ "physics.bio-ph", "cond-mat.soft", "physics.chem-ph" ]
Protein-surface interactions are ubiquitous in biological processes and bioengineering, yet are not fully understood. In biosensors, a key factor determining the sensitivity and thus the performance of the device is the orientation of the ligand molecules on the bioactive device surface. Adsorption studies thus seek to determine how orientation can be influenced by surface preparation. In this work, protein orientation near charged nanosurfaces is obtained under electrostatic effects using the Poisson-Boltzmann equation, in an implicit-solvent model. Sampling the free energy for protein GB1D4' at a range of tilt and rotation angles with respect to the charged surface, we calculated the probability of the protein orientations and observed a dipolar behavior. This result is consistent with published experimental studies and combined Monte Carlo and molecular dynamics simulations using this small protein, validating our method. More relevant to biosensor technology, antibodies such as immunoglobulin G are still a formidable challenge to molecular simulation, due to their large size. We obtained the probability distribution of orientations for the iso-type IgG2a at varying surface charge and salt concentration. This iso-type was not found to have a preferred orientation in previous studies, unlike the iso-type IgG1 whose larger dipole moment was assumed to make it easier to control. We find that the preferred orientation of IgG2a can be favorable for biosensing with positive surface charge of 0.05C/m$^{2}$ or higher and 37mM salt concentration. The results also show that local interactions dominate over dipole moment for this protein. Improving immunoassay sensitivity may thus be assisted by numerical studies using our method (and open-source code), guiding changes to fabrication protocols or protein engineering of ligand molecules to obtain more favorable orientations.
physics.bio-ph
physics
Probing protein orientation near charged nanosurfaces for simulation-assisted biosensor design Christopher D. Cooper,1, 2, a) Natalia C. Clementi,3, b) and Lorena A. Barba3, c) 1)Mechanical Engineering, Boston University, Boston, MA 2)Mechanical Engineering, Universidad T´ecnica Federico Santa Mar´ıa, Valpara´ıso, Chile 3)Mechanical & Aerospace Engineering, The George Washington University, Washington, DC. (Dated: 5 November 2018) Protein-surface interactions are ubiquitous in biological processes and bioengineering, yet are not fully under- stood. In biosensors, a key factor determining the sensitivity and thus the performance of the device is the orientation of the ligand molecules on the bioactive device surface. Adsorption studies thus seek to determine how orientation can be influenced by surface preparation, varying surface charge and ambient salt concen- tration. In this work, protein orientation near charged nanosurfaces is obtained under electrostatic effects using the Poisson-Boltzmann equation, in an implicit-solvent model. Sampling the free energy for protein G B1 D4(cid:48) at a range of tilt and rotation angles with respect to the charged surface, we calculated the prob- ability of the protein orientations and observed a dipolar behavior. This result is consistent with published experimental studies and combined Monte Carlo and molecular dynamics simulations using this small protein, validating our method. More relevant to biosensor technology, antibodies such as immunoglobulin G are still a formidable challenge to molecular simulation, due to their large size. With the Poisson-Boltzmann model, we obtained the probability distribution of orientations for the iso-type IgG2a at varying surface charge and salt concentration. This iso-type was not found to have a preferred orientation in previous studies, unlike the iso-type IgG1 whose larger dipole moment was assumed to make it easier to control. Our results show that the preferred orientation of IgG2a can be favorable for biosensing with positive charge on the surface of 0.05C/m2 or higher and 37mM salt concentration. The results also show that local interactions dominate over dipole moment for this protein. Improving immunoassay sensitivity may thus be assisted by numerical studies using our method (and open-source code), guiding changes to fabrication protocols or protein engineering of ligand molecules to obtain more favorable orientations. I. INTRODUCTION Protein adsorption plays a key role in many biotech- nological applications, particularly biomaterials and tis- implants and biosensors. sue engineering, biomedical Yet, despite their importance, the specific mecha- nisms governing protein-surface interactions are not fully understood.1,2 In the field of biosensors, protein adsorption needs to be engineered to obtain a successful device. Biosensors detect specific molecules using a nanoscale sensing ele- ment, such as a metallic nanoparticle or nanowire cov- ered with a bioactive coating. The prevailing method to modify a sensor surface is via a self-assembled monolayer (sam) of a small charged group, with ligand molecules layered on top to achieve the desired function. Anti- bodies are a common choice for the ligand molecules, although the newest devices use single-domain or single- chain fragment molecules.3,4 Sensing occurs when a tar- get biomolecule binds to the ligand molecule, changing some physical parameter on the sensor, such as current a)[email protected],[email protected] b)[email protected] c)[email protected]; http://lorenabarba.com/ in nanowires or plasmon resonance frequency in metallic nanoparticles. One of the factors crucially affecting biosensor perfor- mance is the orientation of ligand molecules.5,6 These have specific binding sites, which need to be accessi- ble to the target molecule for the biosensor to function well. Probing protein orientation is thus one key goal of adsorption studies. The aim of this study is to as- certain how orientation can be influenced by fabrication conditions regarding surface preparation, such as surface charge and ambient salt content. We consider in particu- lar the antibody immunoglobulin G near a solid surface at different charge concentrations and ionic strengths. Us- ing a smaller molecule (protein G B1 D4(cid:48) ), we could first confirm agreement of our results with published works reporting experiments7 and simulations with a combined Monte Carlo and molecular dynamics method.8 These previous works, among others, also concluded that elec- trostatic interactions are the dominant effect in the orien- tation of adsorbed proteins. In the case of immunoglob- ulin G (IgG), the protein is relevant for biosensor appli- cations, but its large size would make all-atom molecular simulations quite cumbersome and expensive. For this reason, other researchers have studied adsorption of IgG using a coarse-grained model that considers each residue as a sphere (united-residue model),9 finding that elec- trostatics dominates the orientation for higher surface 5 1 0 2 g u A 0 2 ] h p - o i b . s c i s y h p [ 4 v 0 5 1 8 0 . 3 0 5 1 : v i X r a charges and that a positive charge can result in the de- sired "tail-on" placement for the IgG1 iso-type, at low enough salt concentration. Here, we investigate the pre- ferred orientations for the IgG2a variant, which other re- searchers found hard to control. In addition to obtaining the preferred orientation at different conditions of charge and ionic strength, we also take a detailed look at the probability distribution in the parameter space. Our model for protein-surface interactions uses the Poisson-Boltzmann equations in their integral formu- lation, representing the protein geometry as a dielec- tric interface in an implicit solvent. We recently veri- fied the model against an analytical solution valid for spherical geometries and studied its numerical conver- gence in detail.10 Previous studies on protein-surface in- teraction using the Poisson-Boltzmann equation showed that such a model is adequate as long as conforma- tional changes in the protein are slight,11,12 and also that van der Waals effects can be neglected for realistic molecular geometries.13 Conformational changes of the biomolecule can be ignored in this case because bind- ing sites need to remain nearly unmodified during the biosensor fabrication process.5 A continuum framework has been used in the past to study protein orientation,14 but it included ions explicitly. Other studies have used a coarse-grained model of the molecule, represented as a set of spheres,15,16 assigned effective charges at the residue level,9,17 or made approximations to account for pH effects.18,19 The sensor element (functionalized with the sam) is represented in our model as a charged surface that in- teracts electrostatically with the biomolecule. A param- eter sweep of the protein's rotation and tilt angles with respect to the solid surface provides energy landscapes, where the probability of finding the system in a given micro-state depends on the total free energy. The con- tinuum approach can thus provide insights to the condi- tions (surface charge and salt concentration) conducive to a favorable orientation of large proteins, too large for all-atom molecular simulation with today's computing power. It can also represent solid surfaces of any geome- try, and we expect that it may in future assist in the de- sign of better ligand-molecule immobilization techniques for high-sensitivity biosensors. 2 determines the closest a water molecule can get to the protein, and we generate it by rolling a spherical probe of the size of a water molecule around the protein. The dielectric constant inside the protein is low ( = 2 to 4) and there are point charges placed at the atomic loca- tions. The solvent region has the dielectric constant of water  ≈ 80, and we need to account for the presence of salt. This model results in a system of partial differential equations where the Poisson equation describes the elec- trostatic potential inside the protein, and the linearized Poisson-Boltzmann equation applies outside the protein. On the ses, appropriate interface conditions ensure the continuity of the potential and electric displacement. FIG. 1: Sketch of a molecule interacting with a surface: Ω1 is the protein, Ω2 the solvent region, Γ1 is the ses and Γ2 a surface with imposed charge. In this work, we use an extension of the implicit-solvent model to consider the effect of charged surfaces. Such is the case of the setup sketched by Figure 1, which is described mathematically by the following equations: ∇2φ1(r) = −(cid:88) qk 1 ∇2φ2(r) = κ2φ2(r) k δ(r, rk) in solute (Ω1), in solvent (Ω2), φ1 = φ2 1 −2 ∂φ1 ∂n ∂φ2 ∂n = 2 ∂φ2 ∂n = σ0 on interface Γ1, and on surface Γ2 (1) IMPLICIT-SOLVENT MODEL FOR PROTEINS II. NEAR CHARGED SURFACES The implicit-solvent model describes a molecular sys- tem as a set of continuum dielectric regions, and com- putes the mean-field potential using electrostatics. For the case where a protein is dissolved in a solvent, we re- quire two of such regions: inside and outside the protein, interfaced by the solvent-excluded surface (ses). The ses where φi is the electrostatic potential in region Ωi, which has a permittivity i, and σ0 is a prescribed charge on the surface. The surface Γ2 could correspond to a device such as a biosensor. Boundary integral formulation -- We apply Green's second identity to the system of partial-differential equa- tions in (1), and evaluate the resulting equations on Γ1 and Γ2 to obtain the following system of integral equa- tions: ⌦1,✏1,1⌦2,✏2,21nn2, (cid:18) ∂ L (φ1,Γ1) − V Γ1 (cid:19) L (cid:18) ∂ (cid:19) ∂n φ1,Γ1 ∂n − K Γ1 (cid:19) Nq(cid:88) k=0 = 1 1 φ1,Γ1 − K Γ1 Y (φ1,Γ1) + V Γ1 Y φ1,Γ1 Y (φ2,Γ2 ) + V Γ1 Y Y (φ1,Γ1) + 1 2 V Γ2 Y φ1,Γ1 + φ2,Γ2 2 − K Γ2 Y (φ2,Γ2 ) + V Γ2 Y 2 −K Γ2 φ1,Γ1 2 + K Γ1 1 2 (cid:18) ∂ ∂n 3 (2) on Γ1, on Γ1, on Γ2. qk (cid:18) (cid:19) 4πrΓ1 − rk (cid:19) (cid:18) − σ0 2 − σ0 2 = 0 = 0 The function φi,Γj = φi(rΓj ) is the electrostatic potential at a point that approaches the surface Γj from the region Ωi, and K and V are known as the single- and double- layer potentials, correspondingly: III. METHODS A. Discretization and implementation details (cid:73) (cid:73) Γj ∂ ∂n ∂ ∂n L/Y (φi,Γj ) = K Γk (cid:18) ∂ (cid:19) V Γk L/Y φi,Γj = Γj ∂n (cid:2)GL/Y (rΓk , rΓj )(cid:3) φi,Γj dΓ, φi,Γj GL/Y (rΓk , rΓj ) dΓ. (3) In Equation (3), GL and GY are the free-space Green's functions of the Poisson and linearized Poisson- Boltzmann equations, respectively. The single-layer po- tential of a distribution ψ on a surface Γ evaluated at r, V r(ψΓ), can be interpreted as the potential on r due to a charge distribution ψ on Γ. Similarly, K r(ψΓ) can be seen as the potential induced by a double layer of charges (ψ) with opposite sign at Γ. Rearranging terms, we write Equation (2) in matrix form, as follows:  L −V Γ1 L V Γ1 Y 1 2 1 1 2 + K Γ1 2 − K Γ1 −K Γ2 Y Y 1 2 V Γ2 Y 0  (cid:16) 1 (cid:17) −K Γ1 (cid:80)Nq 2 − K Γ2 Y Y ∂ ∂n φ1,Γ1 φ2,Γ2  φ1,Γ1 (cid:17) (cid:16) σ0 (cid:17) (cid:16) σ0 qk 2  =  . 2 V Γ1 Y V Γ2 Y k=0 4πrΓ1−rk  (4) We solve the system in (2) numerically using a bound- ary element method (bem). To represent the ses, we use flat triangular panels where the potential (φ) and its nor- mal derivative (∂φ/∂n) are constant, and then collocate the discretized equation on the center of each panel. This transforms the integral operators in the matrix equation (4) into block matrices of size Np × Np, where Np is the number of panels. Each entry of the block matrix is an integral over one panel (Γj), evaluated on the center of panel Γi: (cid:90) (cid:90) KL,ij = ∂ ∂n Γj (cid:2)GL(rΓi, rΓj )(cid:3) dΓj, VL,ij = GL(rΓi, rΓj )dΓj. (5) Γj We classify the integrals in Equation (5) in three groups, depending on the distance d between the panel and the collocation point. When the collocation point is inside the panel being integrated, we get a singular inte- gral that we solve with a semi-analytical approach28 plac- ing Gauss nodes on the sides of the triangle. We call near- 2 · Area). singular integrals those where d < 2L (L = For near-singular integrals, we use a high-order Gauss quadrature rule with 19 or more nodes. Finally, when the panel and collocation points are further than 2L from each other, we only need 1, 3, 4 or 7 Gauss nodes per el- ement to get good accuracy. √ To solve the resulting linear system, we use a gen- eral minimal residual method (gmres). The most time consuming part of the gmres solver is a matrix-vector multiplication -- in principle, an O(N 2) operation -- done within every iteration of the solver. But by using a treecode algorithm, we perform this operation in O(N log N ) time.29 More details on our implementation of the bem can be found in our earlier work,30 and a companion paper.10 The boundary-integral formulation is not limited to represent the protein with a single surface, but can ac- count for solvent-filled cavities inside the protein region and Stern layers.20 In those cases, more than one surface is required to appropriately represent the protein. Our implementation follows the guidelines from Altman and co-workers21 to deal with multiple surfaces. The boundary-integral formulation of the implicit- solvent model is a popular alternative to compute sol- vation energies of proteins,20 -- 26 but the effect of charged surfaces has rarely been considered. The only work that we know of that does include these effects is limited to plane, infinite surfaces.27 B. Energy calculation We can decompose the total free energy into Coulom- bic, surface, and solvation energy: FTotal = FCoulomb + Fsurf + Fsolv. (6) 4 Coulombic energy -- The Coulombic energy arises simply from the Coulomb interactions of all point charges. We compute it by Nq(cid:88) Nq(cid:88) j i i(cid:54)=j FCoulomb = 1 2 qiqj 1 4πri − rj (7) Solvation free energy -- The solvation energy is the energy contribution of the protein's surroundings: sol- vent polarization, charged surfaces, and other proteins. We compute it as (cid:90) Nq(cid:88) 1 2 k=0 Ω Fsolv = = ρ (φtotal − φCoulomb) qk(φtotal − φCoulomb)(rk), (8) (9) where ρ is the charge distribution, consisting of point charges (which transforms the integral into a sum), and φreac = φtotal − φCoulomb is φreac,rk = −K rk L (φ1,Γ1) + V rk L φ1,Γ1 (10) (cid:19) (cid:18) ∂ ∂n Surface free energy -- We use the description of free energy of a surface with prescribed charge (like Γ2 in Figure 1) from Chan and co-workers.31,32 They describe the free energy on a surface as (cid:90) Γ Fsurf = 1 2 where φ = Gcσ0. FIG. 2: Setup of the problem for our orientation-sampling studies. where Λ is the ensemble of all micro-states, kB is the Boltzmann constant and T the temperature. To obtain a probability distribution, we assume that electrostatic effects are dominant and use Equation (12), sampling Ftotal for different orientations. We define the orientation using the angle between the dipole moment and surface normal vectors as a reference (tilt angle), varying from 0◦ to 180◦. For each tilt angle, we rotate the protein about the dipole moment vector by 360◦ to examine all possible orientations. This process is sketched in Figure 2. In this case, micro-states are defined by the tilt (αtilt) and rotational (αrot) angles, and we rewrite the integral in the numerator of Equation (12) as: (cid:90) (cid:18) exp λ − Ftotal kBT (cid:19) (cid:90) (cid:90) (cid:18) (cid:19) − Ftotal kBT dλ = exp dαrotdαtilt, Gcσ2 0dΓ, (11) (13) where micro-state λ is a range of angles αrot and αtilt. In biosensors, the ligand is adsorbed on the surface (usually covalently), hence we are interested on the orientation of the molecule very close to the surface, and don't consider configurations away from it. C. Orientation sampling of a protein near a charged surface D. Structure preparation We are aiming to investigate the orientation of pro- teins near self-assembled monolayers (sam), specifically for biosensing applications. In the framework of the implicit-solvent model, we can represent the sam as a sur- face charge density, and use Equation (4) to compute the electrostatic potential. According to the Boltzmann dis- tribution, the probability of finding the system in micro- state λ depends on the total free energy, Ftotal, as follows: (cid:82) (cid:82) λ exp Λ exp (cid:16)− Ftotal (cid:16)− Ftotal kB T kB T (cid:17) (cid:17) P (λ) = dλ dΛ , (12) To assess the adequacy of the implicit-solvent model for investigating protein-surface interactions, we stud- ied the orientation of protein G B1 D4(cid:48) mutant near a charged surface, since there are results available in the literature that we could compare to: both experimental observations7 and simulations using a combined Monte Carlo and molecular dynamics approach.8 Figure 3 shows the structure of protein G B1 D4(cid:48) (PDB code 1PGB), to which we applied mutations E19Q, D22N, D46N and D47N to obtain the D4(cid:48) mutant, using the SwissPdb Viewer software.33 We then studied the orientation of antibody Test cases -- span orientation anglesVary tilt angle 0 to 180º, every 2ºVary rotation angle 0 to 360º, every 5º Tilt angleRotation angleDipole momentSurface charge:0.05 C/m2100Å100Å10Å 5 ter, with 32 gpu nodes featuring dual Intel Xeon E2620 2.0GHz 6-core processors with dual nvidia K20 and single-cpu 128 GB of memory. All runs were serial: and single-gpu. We obtained the van der Waals radii and charge distribution using pdb2pqr35 with an amber forcefield, and generated the meshes using the free msms software.36 In these tests, we did not consider a Stern layer for either the protein or the charged surface, nor the presence of solvent-filled cavities inside the protein. A. First case: protein G B1 D4(cid:48) We investigated the preferred orientation of protein placed 2A away from a 100A×100A×10A block G B1 D4(cid:48) with surface charge density ±0.05C/m2, centered with respect to a 100A×100A face. In biosensors, protein G B1 D4(cid:48) can be used as an intermediate protein, coupled to the functionalized surface directly by covalent bond- ing. The protein will thus be at a small distance from the surface. In this case, 2A is in the order of magnitude of the size of a water molecule, or of a C -- N bond. The charge density of ±0.05C/m2 matches that used in other works.8 In these cases, we considered a solvent with no salt, i.e., κ = 0 (to compare with other published results), and with relative permittivity 80. The region inside the protein had a relative permittivity of 4. As seen in Figure 2, αtilt is the angle between the pro- tein's dipole moment and the normal vector to the sur- face, and αrot rotates about the dipole moment. When the dipole-moment vector and the normal are aligned (αtilt = 0), we define a vector Vref as the shortest dis- tance between the axis normal to the surface that goes through the center of mass, and the atom that is furthest away from it. We use Vref as a reference to define the ro- tation angle αrot: the angle between Vref and the vector normal to a 100A×10A face. We sampled the total free energy every ∆αtilt = 2◦ of tilt angle and ∆αrot = 10◦ of rotation angle, resulting in 3, 240 independent runs. The surface mesh had 4 trian- gles per square Angstrom on the protein geometry and 2 triangles per square Angstrom on the charged surface. Numerical parameters are presented in Table I. In a companion publication,10 we present a grid-convergence study using both an analytical solution and a case with protein G B1 D4(cid:48) .37 We computed an approximate ex- act value of −222.43[kcal/mol] for solvation energy and 317.98[kcal/mol] for surface energy using Richardson ex- trapolation with very fine parameters. With results that are less than 2% away from the approximate exact val- ues, we are comfortable with the parameters in Table I and mesh densities of 4 elements per square Angstrom on the protein and 2 elements per square Angstrom on the surface.10 Using total free energy as the input, the integrals of Equation (13) can be computed by means of the trape- zoidal rule. Figure 5 presents the probability of the protein orientation in terms of cos(αtilt), in intervals of FIG. 3: Structure of protein G B1 D4(cid:48) (PDB code 1PGB). FIG. 4: Structure of immunoglobulin G (PDB code 1IGT). immunoglobulin G iso-type IgG2a (PDB code 1IGT), a widely used protein in biosensors, whose structure is shown in Figure 4. This is a more interesting case from the point of view of our application, yet it is more diffi- cult to study with molecular simulation due to its size. In both cases, the vector orientation of the dipole moment (used as reference for the tilt and orientation angles) was obtained using the location of the point charges at the locations of the atoms. IV. RESULTS The results detailed in this section were obtained using an extension of the open-source code PyGBe,34 account- ing for the presence of surfaces with imposed charge or potential.10 We ran the calculations for protein G B1 D4(cid:48) on a workstation with Intel Xeon X5650 cpus and one nvidia Tesla C2075 gpu card (late 2011). The second case considers the antibody immunoglobulin G, which is a much larger molecule than protein G B1 D4(cid:48) . For these runs, we used either: (1) Boston University's bungee cluster, which has 16 nodes with 8 Intel Xeon cpu cores each, and a total of 3 nvidia Tesla K20 (Kepler, late 2012) and 26 nvidia Tesla M2070/2075 gpus; or (2) the George Washington University's Colonial One clus- TABLE I: Numerical parameters used for numerically probing the orientation of protein G B1 D4(cid:48) . each run, there is no guarantee that the refinement is ho- mogeneous throughout the whole molecular surface. The numerical parameters are presented in Table III. 6 # Gauss points: in-element close-by far-away Ncrit P θ 9 per side tol. 300 4 0.5 10−5 Treecode: gmres: 19 1 ∆ cos(αtilt) = 0.005 (Fig. 5a) and ∆αtilt=2◦ (Fig. 5b). Table II presents the average orientation < cos(αtilt) > for the surface having either positive or negative charge density, and Figure 6 shows the electrostatic potential for the preferred orientation in each case. TABLE II: Average orientation. < cos(αtilt) > Negative Positive −0.968 0.963 B. Second case: immunoglobulin G We computed the electrostatic field of immunoglobulin G -- a protein widely used in biosensors -- interacting with a 250A×250A×10A block, varying the conditions of sur- face charge and salt concentration. The protein was cen- tered with respect to a 250A×250A face, at a distance 5A above it. In fabrication, antibodies are usually immobi- lized on the biosensor surface via a cross-linker molecule, which we model here by increasing the distance from the surface. As before, the solvent had relative permittivity of 80 and the protein of 4. for study Grid-convergence immunoglobulin G -- Since this was the first time we did calculations on im- munoglobulin G, we carried out a grid-convergence study to make sure the geometry was well resolved and to find adequate values of the simulation parameters for sam- pling different orientations. The error plotted in Figure 7 is the relative difference between the energy obtained using PyGBe with each mesh density and the estimated exact value computed with Richardson extrapolation. In this case, we computed the solvation energy and surface energy of a system consisting of a surface with charge density 0.05C/m2 and a protein with αtilt = 31◦ and αrot = 130◦. Using the results from runs with a mesh density of 2, 4, and 8 elements per square Angstrom, we added the solvation and surface energies, and used Richardson extrapolation to obtain a value of −2792.22[kcal/mol], and an observed order of conver- gence of 0.85. This is our reference to calculate the errors in Figure 7. There is a slight deviation from the expected value of the observed order of convergence (1.0), which we attribute to the non-uniform mesh generated by msms. Even though the mesh density is on average doubled for TABLE III: Numerical parameters used in the grid-convergence study with immunoglobulin G. # Gauss points: in-element close-by far-away Ncrit P θ 9 per side tol. 1000 6 0.5 10−5 Treecode: gmres: 19 1 Probing orientation of immunoglobulin G -- We sam- pled the total free energy every ∆αtilt = 4◦ of tilt angle and ∆αrot = 20◦ of rotation angle, resulting in a to- tal of 810 runs. The surface meshes had 2 triangles per square Angstrom throughout. Numerical parameters are presented in Table IV. TABLE IV: Numerical parameters used in the runs probing orientation of immunoglobulin G. # Gauss points: in-element close-by far-away Ncrit P θ 9 per side tol. 300 2 0.5 10−4 Treecode: gmres: 19 1 With the computed total free energy, we obtained the probability of each orientation using Equation (13) and the trapezoidal rule. We sampled all combinations with surface charges of σ = ±0.05C/m2 and σ = ± 0.1C/m2 and salt concentrations of 145mM (κ = 0.125 A−1) and 37mM (κ = 0.0625 A−1). For each of these cases, Figures 8 and 9 show a color plot of the probability distribution with respect to the tilt and rotation angles, and a 3D plot of the preferred orientation, where the solvent-excluded surface is colored by the electrostatic potential. C. Reproducibility and data management We have a consistent reproducibility practice that in- cludes releasing code and data associated with a pub- lication. The PyGBe code was released at the time of submitting our previous publication,20 under an MIT open-source license, and we maintain a version-control repository. As with our previous paper, we also release with this work all of the data needed to run the numeri- cal experiments reported here, including running scripts and post-processing code in Python for producing the figures. To support our open-science goals, we prepared such a "reproducibility package" for each of the results presented in Figures 5, 7, and the probability plots in Figures 8 and 9. The included running scripts invoke the PyGBe code with the correct input data and meshes (also included), and post-process the results to give the final figure, all with just one command. Please see the respective captions for a reference to the reproducibility packages, hosted on the figshare repository. 7 (a) (b) (c) (d) FIG. 5: Orientation probability distribution of protein G B1 D4(cid:48) . Figures 5a and 5b are the probability with respect to the tilt angle and its cosine, respectively. Figures 5c and 5d are the probability distributions with respect to both the tilt and rotation angles. Data sets, figure files and running/plotting scripts are available under cc-by.38 (a) Negative surface charge (αtilt = 172◦, αrot = 110◦) FIG. 6: Electrostatic potential of protein G B1 D4(cid:48) indicates direction of dipole-moment vector. (b) Positive surface charge (αtilt = 8◦, αrot = 150◦) for the preferred orientations according to Figure 5. Black arrow V. DISCUSSION A. First case: protein G B1 D4(cid:48) The orientation of protein G B1 D4(cid:48) near charged sur- faces was studied using a combined Monte Carlo and molecular dynamics method by Liu and co-workers8 and experimentally by Baio and co-workers.7 The availability of these published results was a motivation to use this protein for a first test, to compare with the results ob- tained with our model. The results presented in Figure 5 show that for the 1.00.50.00.51.0cos(αtilt)0.000.050.100.150.200.250.300.35Probability›cos(αtilt)fiσ=0.05=0.963›cos(αtilt)fiσ=−0.05=-0.968σ=0.05 C/m2σ=-0.05 C/m204590135180αtilt0.000.020.040.060.080.100.12Probability›αtiltfiσ=0.05=13.2◦›αtiltfiσ=−0.05=167.7◦σ=0.05 C/m2σ=-0.05 C/m204590135180αtilt050100150200250300350αrotProbability for σ=0.050.00000.00040.00080.00120.00160.00200.00240.00280.00320.003604590135180αtilt050100150200250300350αrotProbability for σ=-0.050.00000.00040.00080.00120.00160.00200.00240.00280.00320.0036N-terminalC-terminalN-terminalC-terminal 8 effects play a role only in cases of very low surface charge. For example, in Ref. 9, van der Waals effects were of consequence in a setup with surface charge of 0.006C/m2 and high ionic strength, leading to weak electrostatics. In a biosensor-fabrication scenario, this would only be the case with low-quality sams. The results with protein G B1 D4(cid:48) mean that an elec- trostatic solver with implicit solvent using the Poisson- Boltzmann equation is capable of capturing the driving mechanism of physical adsorption and orientation of the adsorbed molecule, at least in cases where the molecule's dipole moment is dominating the orientation. This is im- portant because protein adsorption, being a free energy- driven process, is difficult to study experimentally41 and thus simulations offer a promising alternative. Full atomistic molecular dynamics, however, demands large amounts of computing effort, and the possibility of using an electrostatics solver may extend the range of systems that can be investigated. B. Second case: immunoglobulin G With our numerical model already verified using an analytical solution for spherical geometry10 and the suc- cessful results for protein orientation of a small protein near a charged surface (Section V A), we proceeded to study the effect of surface charge and salt concentra- tion on the orientation of the antibody immunoglobu- lin G. Antibodies are widely used in biosensors as lig- and molecules, due to their affinity and specificity with the target molecule (antigen), and it is vitally important that they are adsorbed on the sensor with the antigen- binding Ig fragment (Fab) pointing away from the sensor, into the oncoming flow containing the antigens (known as "end-on" or "tail-on" orientation). Early experimen- tal studies found that antigen/antibody ratio was espe- cially low on negatively charged surfaces,42 leading to the notion that protein orientation was affected to leading order by charge. One subsequent study43 investigated the orientations of two iso-types of immunoglobulin G -- IgG1, corresponding to PDB structure 1IGY, and IgG2a, corresponding to PDB 1IGT -- adsorbed on positive and negatively charged surfaces. As an indirect method of probing antibody orientation, the researchers obtained adsorbed amounts and antigen/antibody ratios by means of surface-plasmon resonance experiments (e.g., a higher antigen/antibody ratio would indicate that more active sites are accessible and more antibodies are in a favor- able orientation). The finding was that IgG1 mainly had a "head-on" (unfavorable) orientation on the negatively charged surfaces and a mix of "tail-on" (most favorable) and "side-on" orientations on the positively charged sur- faces. IgG2a, on the other hand, had many orientations on both surfaces with positive and negative charge, lead- ing to the conclusion that IgG2a is harder to control using electrostatic effects. Results consistent with these were obtained by Zhou and co-workers9 using a united-residue FIG. 7: Grid-convergence study of the solvation plus surface energy for immunoglobulin G interacting with a surface with charge density of 0.05C/m2. Data sets, figure files and plotting scripts are available under cc-by.39 most likely orientations, the dipole-moment vector is aligned with the vector normal to the interacting surface. This indicates that the dipole moment is the dominant effect that determines the protein's orientation, over local protein-surface interactions. This is the expected result, since protein G B1 D4(cid:48) is a relatively small biomolecule. Moreover, Figure 5 reveals that protein G B1 D4(cid:48) be- haves like a point dipole, as the most likely orientations shift 180◦ when the sign of the surface charge is flipped. This is also explained by the dipole moment dominat- ing the orientation. In fact, we repeated this whole set of calculations but placing protein G B1 D4(cid:48) at a greater distance, 5A away from the surface, and the results did not vary. adsorbed on NH+ The dipolar behavior described by our results agrees with the experiments done by Baio and co-workers,7 in which they observed opposite orientations of protein 3 and COO− self-assembled G B1 D4(cid:48) monolayers. With positively charged surfaces, most of the proteins oriented with the N-terminal of the pro- tein pointing away from the surface, while for nega- tively charged surfaces the opposite occurred, with the C-terminal pointing away from the surface. This agrees with our results in Figure 5 (the dipole moment vector of protein G B1 D4(cid:48) points from the C-terminal to the N- terminal). Liu and co-workers8 used a combined Monte Carlo and molecular dynamics method to obtain < cos(αtilt) >= 0.95 for σ = 0.05C/m2, and < cos(αtilt) >= −0.85± 0.05 for σ = −0.05C/m2, which agrees well with our re- sults in Table II. Note that MD simulations consider van der Waals interactions and conformational changes of the protein, whereas these are not considered in our ap- proach, explaining the slight differences in < cos(αtilt) >. However, as noted by other researchers,7 -- 9 electrostatic effects often dominate protein-surface interactions and drive orientation during adsorption, while van der Waals 100101Mesh density10-210-1Relative errorN−1 9 (a) Probability for σ = −0.05C/m2, κ = 0.125A−1 (b) x-y plane view for αtilt = 116◦ and αrot = 100◦ (c) Probability for σ = −0.1C/m2, κ = 0.125A−1 (d) y-z plane view for αtilt = 36◦ and αrot = 300◦ (e) Probability for σ = −0.05C/m2, κ = 0.0625A−1 (f) x-y plane view for αtilt = 40◦ and αrot = 340◦ (g) Probability for σ = −0.1C/m2, κ = 0.0625A−1 (h) x-y plane view for αtilt = 40◦ and αrot = 40◦ FIG. 8: Orientation probability distribution and surface potential of the preferred orientation for immunoglobulin G near a negative surface charge. The black arrow indicates the direction of the dipole moment, and the circles enclose the Fab fragments. Data sets, figure files and plotting scripts available under cc-by.40 04590135180αtilt050100150200250300350αrotProbability for σ=-0.050.000.010.020.030.040.050.060.070.080.09xzy04590135180αtilt050100150200250300350αrotProbability for σ=-0.100.0000.0080.0160.0240.0320.0400.0480.0560.0640.072xzy04590135180αtilt050100150200250300350αrotProbability for σ=-0.050.0000.0150.0300.0450.0600.0750.0900.105yxz04590135180αtilt050100150200250300350αrotProbability for σ=-0.100.000.020.040.060.080.100.120.140.160.18yxz 10 (a) Probability for σ=0.05C/m2 and κ=0.125A−1 (b) x-y plane view for αtilt = 64◦ and αrot = 280◦ (c) Probability for σ=0.1C/m2 and κ=0.125A−1 (d) y-z plane view for αtilt = 32◦ and αrot = 100◦ (e) Probability for σ=0.05C/m2 and κ=0.0625A−1 (f) y-z plane view for αtilt = 44◦ and αrot = 120◦ (g) Probability for σ=0.1C/m2 and κ=0.0625A−1 (h) x-y plane view for αtilt = 64◦ and αrot = 260◦ FIG. 9: Orientation probability distribution and surface potential of the preferred orientation for immunoglobulin G near a positive surface charge. The black arrow indicates the direction of the dipole moment, and the circles enclose the Fab fragments. Data sets, figure files and plotting scripts available under cc-by.40 04590135180αtilt050100150200250300350αrotProbability for σ=0.050.000.010.020.030.040.050.060.070.08yxz04590135180αtilt050100150200250300350αrotProbability for σ=0.100.0000.0150.0300.0450.0600.0750.0900.1050.120yxz04590135180αtilt050100150200250300350αrotProbability for σ=0.050.000.030.060.090.120.150.180.210.240.27yxz04590135180αtilt050100150200250300350αrotProbability for σ=0.100.000.040.080.120.160.200.240.28yxz model: a coarse-grained model where each amino-acid is treated as a sphere. They find that IgG1 will have the fa- vorable "end-on" orientation on positive surfaces, as long as the charge density was large enough (0.018C/m2, in their case) and the ionic strength was low. But IgG2a did not show a clear preferred orientation at the conditions they looked at; the authors attribute this to the weaker dipole moment of this iso-type. We investigated the orientation of IgG2a, which other studies found harder to orient favorably on a biosen- sor surface, and used two values of the surface charge (σ = 0.05 and 0.1C/m2) and two values of salt concen- tration (κ = 0.125 and 0.0625A−1), in each case varying two-fold. Figures 8 and 9 present the probability dis- tribution of IgG2a for many orientations (given by αtilt and αrot), in each case. The following discussion refers to each variation of the parameters and the effect on the preferred orientation of the adsorbed antibody and its probability. Effect of surface charge -- The lower value of sur- face charge here is σ = ±0.05C/m2, the same value used in Ref. 8 to mimic the experiments reported in Ref. 7. Figures 8a and 9a show that for the lower value of surface charge with the higher salt concentra- tion (κ = 0.125A−1), there is no clear preferred orienta- tion, to the point that the highest probability falls under 10%. This means that adsorbing the antibodies under these conditions would result in a wide range of orienta- tions, which would not be favorable for biosensor fabri- cation. Moreover, the preferred configurations in figures 8a and 9a show the antibody lying flat on the surface, far from the desired "tail on" orientation. This observa- tion is consistent with a previous study using a unified- residue model,9 where this particular antibody showed many possible orientations. The authors of that study attributed this behavior to the weaker dipole moment of this molecule, compared with the variant IgG1. With the higher value of surface charge, in this case σ = ±0.1C/m2, the orientation probability distribution in the case of negative charge improves somewhat, as the antibody is slanted sideways rather than lying down for κ = 0.125A−1 (at least one antingen-binding fragment is pointing up), and the probability of the preferred ori- entation is almost doubled for low salt concentration, in a "side on" orientation. For positive surface charge the slanted orientation is similar, however the probability is higher for the preferred orientation in both the low- and high-salt cases. In the cases with higher salt concen- tration, Figure 9c shows a preferred orientation with a higher probability of 12%, compared to 8% in Figure 9a, and the dipole moment rotates towards the normal vec- tor. For the lower value of salt concentration (Figure 9g), this effect is smaller, however it shifts the preferred tilt angle in the opposite direction, from 44◦ to 64◦. Note that the dipole-moment vector does not point straight through the middle between the two Fab fragments, but in an angle. This indicates that, in contrast to protein G B1 D4(cid:48) , local interactions dominate over the dipole mo- 11 ment. If the dipole moment were the dominant effect, the dipole-moment vector would tend to align to the sur- face normal as the surface charge increases. This argues against the suggestion by other researchers9,43 that the dipole-moment vector is the main determinant of orien- tation. Effect of salt concentration -- As the surface charge density was varied two-fold, we also varied the Debye length (κ−1) two-fold. In terms of salt concentration, it means a 4× decrease in the amount of salt. The higher value of salt concentration corresponds to 145mM, which is in the physiological salt range. Like increasing the surface charge, lowering the salt concentration affects the orientation probability distri- bution. For σ = −0.05C/m2 (Fig. 8e), the effect is a large shift in the preferred tilt angle, from αtilt = 116◦ to αtilt = 40◦, with a small change in probability. For the positive weaker charge, σ = 0.05C/m2 (Fig. 9e), not only does the peak probability increase considerably (∼ 3×), but the preferred tilt shifts from 64◦ to 44◦. This ori- entation is favorable for biosensing applications, as the antigen-binding fragments are pointing away from the surface, in a "tail on" orientation. For the stronger neg- ative charge, σ = −0.1C/m2, the probability peak in- creases 2.5× for a "side on" orientation where one of the antigen-binding fragments is attached to the surface in an unfavorable position (Fig. 8g). With positive surface charge, the tilt angle shifts in such a way that the anti- body is lying on the surface with a marked probability close to 30% (Fig. 9g). This orientation is not ideal for biosensors, but it is better than the slanted position as neither of the Fabs are attached to the surface. From the results in Figures 8 and 9, we conclude that the iso-type IgG2a can in general be better orientated with low salt concentration and high surface charge, as we get more pronounced high-probability regions. Moreover, good orientations for biosensors are more likely to occur with positive surface charge (Figure 9f), since the Fab fragments are pointing up. Previous studies had shown that the IgG1 variant could be controlled, but not IgG2a. The advantage of a positive surface charge and a low ionic strength had been suggested by previous studies, but not for this particular variant of immunoglobulin G. Note also that our lower value of salt concentration is 37mM, which is a higher amount of salt than other studies.42,43 A limitation of this study stems from the appli- cation of linearized Poisson-Boltzmann equation. Bu and co-workers44 assessed the accuracy of the Poisson- Boltzmann equation for highly charged surfaces (∼ 0.4C/m2), getting good agreement of the model with ex- periments. Rigorously, the linearized Poisson-Boltzmann equation is a valid approximation of the Poisson- Boltzmann equation when the nondimensional potential is smaller than 1 (φqe/kBT << 1). However, this restric- tion can be relaxed when calculating solvation energy. For example, we ran a calculation using our boundary element code on an isolated IgG2a immersed in a solvent with 37mM of salt (κ = 0.0625A−1), with the parameters from Table IV. Computing the absolute value of the di- mensionless potential on the molecular surface gives over 55% of the triangles with φqe/kBT > 1 and an average value of 1.5. Yet, using the linear or non-linear Poisson- Boltzmann model of APBS,45 the solvation energy for the isolated IgG2a gives the same result. (The APBS tests used a volumetric mesh of 150 × 150 × 150A3 with 4493 nodes.) Adding a surface with charge σ = 0.1C/m2 next to IgG2a in the configuration of Fig. 9h, the situa- tion is similar: 51% of the triangles have a dimensionless potential with absolute value exceeding 1 and the aver- age is 1.3. By comparison with the isolated IgG2 case, we expect linearized Poisson Boltzmann to give a good approximation of the solvation energy in this case. This conclusion is also consistent with the results of Ref. 46 that show good agreement between linear and non-linear Poisson Boltzmann when the average dimensionless po- tential on the molecular surface is between −2 and 2. Finally, with a value of surface charge of 0.1C/m2, the mechanism of protein orientation appears to be dictated by the solvation energy, rather than the surface energy. That is, the maximum probability of the preferred orien- tation occurs at the minimum of solvation energy.47 Since solvation energy in our systems is well approximated by linear PB, we conclude that the use of this theory is jus- tified in these cases. Our results suggest that a combination of high surface charge and low salt concentration increases the proba- bility of the preferred configuration, and that more fa- vorable orientations are obtained with positive surface charge. We completed an additional set of tests for the orientation of IgG2a near a surface with σ = 0.2C/m2 and κ = 0.03125A−1, even though the electrostatic po- tential is outside the linear regime in this case. The result in Figure 10 suggests that as we increase the electrostatic effects, the preferred orientation becomes highly marked, and tends towards a favorable orientation for biosensors. However, our model is based on the linearized Poisson- Boltzmann equation and this test goes outside the known regime where the model can be applied. We cannot claim that this is a physical situation, but it indicates a trend. Further study of the conditions that make IgG2a orient favorably for biosensors may need nonlinear models, and combined experimental observations. VI. CONCLUSION Various studies have revealed the importance of pro- tein orientation in immunoassays. One work suggested that highly oriented antibodies could result in 100× im- provement in the affinity of a biosensor.5 Thus, a design goal would be to know how to prepare a surface to control protein orientation. Yet, despite much work, control of protein orientation has not been successful. This study increases our understanding of how nanosurface proper- ties (charge) and preparation conditions (salt levels) af- 12 fect protein orientation. We successfully used an implicit- solvent model to study protein orientation near charged surfaces, which in our method can have any geometry. In a companion publication,10 we describe expanding the applicability of our open-source code, PyGBe, to account for the presence of charged surfaces and present grid- convergence studies using an analytical solution and pro- tein G B1 D4(cid:48) . Protein G B1 D4(cid:48) behaves like a point dipole near a charged surface, with the dipole-moment vector shifting ∼180◦ when the sign of the surface charge flips. Our re- sults compare well with experimental observations and simulations using combined Monte Carlo and molecu- lar dynamics methods, supporting the use of our ap- proach for probing protein orientation near charged sur- faces. We applied our approach to immunoglobulin G, a biomolecule that is much larger than protein G B1 D4(cid:48) (about 125×, by volume) and would be challenging to study via molecular dynamics. The iso-type IgG2a was found by previous studies to be hard to control, exhibit- ing many orientations, but we are able to obtain a pre- ferred orientation that is favorable for biosensing with a positive surface of 0.05C/m2 or higher d 37mM of salt in the solvent. We conclude that local electrostatic in- teractions dominate over the dipole moment, and even this protein can be favorably oriented with the appropri- ate fabrication protocol. Potentially, protein engineering could be used to obtain ligand molecules that interact with charged surfaces in a desired fashion. In this appli- cation, where ligand molecules undergo little conforma- tional change as they adsorb on the sensor surface, our new implicit-solvent model can offer a valuable approach to assist in biosensor design. In our future work, and in collaboration with experimental researchers, we intend to use this approach to aid the design of better ligand molecules, by looking at the preferred orientations for different ligand mutants. ACKNOWLEDGMENTS This work was supported by ONR via grant #N00014- 11-1-0356 of the Applied Computational Analysis Pro- gram. LAB also acknowledges support from NSF CA- REER award OCI-1149784 and from NVIDIA, Inc. via the CUDA Fellows Program. We are grateful for many helpful conversations with members of the Materials and Sensors Branch of the Naval Research Laboratory, espe- cially Dr. Jeff M. Byers and Dr. Marc Raphael. 1J. J. Gray, "The interaction of proteins with solid surfaces," Curr. Opin. Struct. Biol. 14, 110 -- 115 (2004). 2M. Rabe, D. Verdes, and S. Seeger, "Understanding protein adsorption phenomena at solid surfaces," Adv. Colloid Interface Sci. 162, 87 -- 106 (2011). 3J.-Y. Byun, Y.-B. Shin, T. Li, J.-H. Park, D.-M. Kim, D.- H. Choi, and M.-G. Kim, "The use of an engineered single chain variable fragment in a localized surface plasmon resonance method for analysis of the c-reactive protein," Chem. Commun. 49, 9497 -- 9499 (2013). 13 (a) Probability distribution (b) x-y plane view for αtilt = 76◦ and αrot = 160◦ FIG. 10: Orientation probability distribution and surface potential of the preferred orientation for immunoglobulin G near a surface with σ = 0.2C/m2 and κ = 0.03125A−1. Note that these conditions are outside the range of linearized theory (as explained in the Discussion). 4A. K. Trilling, T. Hesselink, A. van Houwelingen, J. H. G. Cordewener, M. A. Jongsma, S. Schoffelen, J. C. M. van Hest, H. Zuilhof, and J. Beekwilder, "Orientation of llama antibodies strongly increases sensitivity of biosensors," Biosen. Bioelectron. 60, 130 -- 136 (2014). 5N. Tajima, M. Takai, and K. Ishihara, "Significance of anti- body orientation unraveled: Well-oriented antibodies recorded high binding affinity," Anal. Chem. 83, 1969 -- 1976 (2011). 6A. K. Trilling, J. Beekwilder, and H. Zuilhof, "Antibody orien- tation on biosensor surfaces: a minireview," Analyst 138, 1619 -- 1627 (2013). 7J. E. Baio, T. Weidner, L. Baugh, L. J. Gamble, P. S. Stayton, and D. G. Castner, "Probing the orientation of electrostatically immobilized protein G B1 by time-of-flight secondary ion spec- trometry, sum frequency generation, and near-edge X-ray ad- sorption fine structure spectroscopy," Langmuir 28, 2107 -- 2112 (2012). 8J. Liu, C. Liao, and J. Zhou, "Multiscale simulations of protein G B1 adsorbed on charged self-assembled monolayers," Langmuir 29, 11366 -- 11374 (2013). 9J. Zhou, S. Chen, and S. Jiang, "Orientation of adsorbed anti- bodies on charged surfaces by computer simulation based on a united-residue model," Langmuir 19, 3472 -- 3478 (2003). 10C. D. Cooper and L. A. Barba, "Poisson-Boltzmann model for protein-surface interactions and grid-convergence study using the PyGBe code," (2015), submitted, preprint on arXiv:1506.03745. 11Y. Yao and A. M. Lenhoff, "Electrostatic contributions to pro- tein retention in ion-exchange chromatography. 1. cytochrome c variants," Anal. Chem. 76, 6743 -- 6752 (2004). 12Y. Yao and A. M. Lenhoff, "Electrostatic contributions to pro- tein retention in ion-exchange chromatography. 2. proteins with vaious degrees of structural differences." Anal. Chem. 77, 2157 -- 2165 (2005). 13C. M. Roth, B. L. Neal, and A. M. Lenhoff, "Van del Waals interactions involving proteins," Biophys. J. 70, 977 -- 987 (1996). 14A. H. Juffer, P. Argos, and J. De Vlieg, "Adsorption of proteins onto charged surfaces: a Monte Carlo approach with explicit ions," J. Comp. Chem. 17, 1783 -- 1803 (1996). 15Y.-J. Sheng, H.-K. Tsao, J. Zhou, and S. Jiang, "Orientation of a Y-shaped biomolecule adsorbed on a charged surface," Phys. Rev. E 66, 011911 (2002). 16J. Zhou, H.-K. Tsao, Y.-J. Sheng, and S. Jiang, "Monte Carlo simulations of antibody adsorption and orientation on charged surfaces," J. Chem. Phys. 121, 1050 -- 1057 (2004). 17A. S. Freed and S. M. Cramer, "Protein-surface interaction maps for ion-exchange chromatography," Langmuir 27, 3561 -- 3568 (2011). 18P. M. Biesheuvel, M. van der Veen, and W. Norde, "A mod- ified Poisson-Boltzmann model including charge regulation for the adsorption of ionizable polyelectrolytes to charged interfaces, applied to lysozyme adsorption on silica," J. Phys. Chem. B 109, 4172 -- 4180 (2005). 19R. A. Hartvig, M. van de Weert, J. Ostergaard, L. Jorgensen, and H. Jensen, "Protein adsorption at charged surfaces: The role of electrostatic interactions and interfacial charge regulation," Langmuir 27, 2634 -- 2643 (2011). 20C. D. Cooper, J. P. Bardhan, and L. A. Barba, "A biomolec- ular electrostatics solver using Python, GPUs and boundary elements that can handle solvent-filled cavities and Stern lay- ers." Comput. Phys. Commun. 185, 720 -- 729 (2014), preprint on arXiv:/1309.4018. 21M. D. Altman, J. P. Bardhan, J. K. White, and B. Tidor, "Ac- curate solution of multi-region continuum electrostatic problems using the linearized Poisson -- Boltzmann equation and curved boundary elements," J. Comput. Chem. 30, 132 -- 153 (2009). 22B. J. Yoon and A. M. Lenhoff, "A boundary element method for molecular electrostatics with electrolyte effects," J. Comput. Chem. 11, 1080 -- 1086 (1990). 23A. H. Juffer, E. F. F. Botta, B. A. M. Vankeulen, A. Vander- ploeg, and H. J. C. Berendsen, "The electric potential of a macromolecule in a solvent: A fundamental approach," Journal of Computational Physics 97, 144 -- 171 (1991). 24B. Lu, X. Cheng, J. Huang, and J. A. McCammon, "Order N al- gorithm for computation of electrostatic interactions in biomolec- ular systems," P. Natl. Acad. Sci. USA 103, 19314 -- 19319 (2006). 25C. Bajaj, S. C. Chen, and A. Rand, "An efficient higher-order fast multipole boundary element solution for Poisson-Boltzmann- based molecular electrostatics," SIAM J. Sci. Comput. 33, 826 -- 04590135180αtilt050100150200250300350αrotProbability for σ=0.200.00.10.20.30.40.50.60.70.80.91.0 848 (2011). 26W. H. Geng and R. Krasny, "A treecode-accelerated boundary integral Poisson-Boltzmann solver for solvated biomolecules," J. Comp. Phys. 247, 62 -- 78 (2013). 27B. J. Yoon and A. Lenhoff, "Computation of the electrostatic interaction energy between a protein and a charged surface," J. Phys. Chem. 96, 3130 -- 3134 (1992). 28Z. Zhu, J. Huang, B. Song, and J. White, "Improving the robustness of a surface integral formulation for wideband impedance extraction of 3D structures," in Proceedings of the 2001 IEEE/ACM Int. Conf. on Computer-Aided Design (2001) pp. 592 -- 597. 29J. Barnes and P. Hut, "A hierarchical O(N log N ) force- calculation algorithm," Nature 324, 446 -- 449 (1986). 30C. D. Cooper and L. A. Barba, "Validation of the PyGBe code for Poisson-Boltzmann equation with boundary element methods," Technical Report on figshare, CC-BY license (2013). 31D. Y. Chan and D. J. Mitchell, "The free energy of an electrical double layer," J. of Coll. and Inter. Science 95, 193 -- 197 (1983). 32S. L. Carnie and D. Y. Chan, "Interaction free energy between identical spherical colloidal particles: The linearized poisson- boltzmann theory," J. of Coll. and Inter. Science 155, 297 -- 312 (1993). 33N. Guex and M. C. Peitsch, "SWISS-MODEL and the Swiss- PdbViewer: An environment for comparative protein modeling," Electrophoresis 18, 2714 -- 2723 (1997), http://www.expasy.org/ spdbv/. 34https://github.com/barbagroup/pygbe. 35T. J. Dolinsky, J. E. Nielsen, J. A. McCammon, and N. A. Baker, "PDB2PQR: an automated pipeline for the setup of Poisson -- Boltzmann electrostatics calculations," Nucleic Acids Research 32, W665 -- W667 (2004). 36M. F. Sanner, A. J. Olson, and J.-C. Spehner, "Fast and ro- bust computation of molecular surfaces," in Proceedings of the eleventh annual symposium on Computational geometry (ACM, 1995) pp. 406 -- 407. 14 37C. D. Cooper and L. A. Barba, "Grid convergence of PyGBe with protein G B1 D4," Data, figures and plottings script on figshare, CC-BY license, http://dx.doi.org/10.6084/m9.figshare. 1348803 (2015). 38C. D. Cooper and L. A. Barba, "Protein orientation near a charged surface using PyGBe and protein G B1 D4," Data, figures and plottings script on figshare, CC-BY license, http://dx.doi.org/10.6084/m9.figshare. 1348804 (2015). 39C. D. Cooper and L. A. Barba, "Grid convergence of PyGBe with immunoglobulin G near a charged surface," Data, figures and plottings script on figshare, CC-BY license, http://dx.doi.org/10.6084/m9.figshare. 1348801 (2015). 40C. D. Cooper and L. A. Barba, "Protein orientation near charged surface using PyGBe with immunoglobulin G," Data, figures and plottings script on figshare, CC-BY license, http://dx.doi.org/10.6084/m9.figshare. 1348802 (2015). 41M. Mijajlovic, M. J. Penna, and M. J. Biggs, "Free energy of adsorption for a peptide at a liquid/solid interface via nonequi- librium molecular dynamics," Langmuir 29, 2919 -- 2926 (2013), pMID: 23394469, http://dx.doi.org/10.1021/la3047966. 42J. Buijs, D. D. White, and W. Norde, "The effect of adsorp- tion on the antigen binding by IgG and its F(ab) 2 fragments," Colloids and Surfaces B: Biointerfaces 8, 239 -- 249 (1997). 43S. Chen, L. Liu, J. Zhou, and S. Jiang, "Controlling antibody orientation on charged self assembled-monolayers," Langmuir 19, 2859 -- 2864 (2003). 44W. Bu, D. Vaknin, and A. Travesset, "How accurate is poisson- boltzmann theory for monovalent ions near highly charged inter- faces?" Langmuir 22, 5673 -- 5681 (2006). 45N. A. Baker, D. Sept, M. J. Holst, and J. A. McCammon, "Elec- trostatics of nanoysystems: Application to microtubules and the ribosome," P. Natl. Acad. Sci. USA 98, 10037 -- 10041 (2001). 46F. Fogolari, P. Zuccato, G. Esposito, and P. Viglino, "Biomolec- ular electrostatics with the linearized poisson-boltzmann equa- tion," Biophys. J. 76, 1 -- 16 (1999). 47See Supplementary Materials.
1109.5389
1
1109
2011-09-25T19:43:54
Water drives peptide conformational transitions
[ "physics.bio-ph" ]
Transitions between metastable conformations of a dipeptide are investigated using classical molecular dynamics simulation with explicit water molecules. The distribution of the surrounding water at different moments before the transitions and the dynamical correlations of water with the peptide's configurational motions indicate that water is the main driving force of the conformational changes.
physics.bio-ph
physics
Water drives peptide conformational transitions Non-linearity and Complexity Research Group, Aston University, Birmingham, B4 7ET, UK Dmitry Nerukh∗ Sergey Karabasov† Whittle Laboratory Cambridge University Engineering Department, 1 JJ Thompson Avenue, Cambridge, CB3 0DY, UK Transitions between metastable conformations of a dipeptide are investigated using classical molecular dynamics simulation with explicit water molecules. The distribution of the surrounding water at different moments before the transitions and the dynamical correlations of water with the peptide's configurational motions indicate that water is the main driving force of the conformational changes. Investigations of protein dynamics have recently led to believe that water plays the major role in protein mo- tion. There is a large body of experimental and sim- ulation evidences [1 -- 4] showing a close connection be- tween the water dynamics and the protein conformations. Frauenfelder and colleagues have experimentally shown that protein dominant conformational motions are slaved by the hydration shell and the bulk solvent [5], while the protein molecule itself provides an 'active matrix' nec- essary for guiding the water's dynamics towards biolog- ically relevant conformational changes. The change in water dynamics at the shell of up to almost a dozen wa- ter molecule diameters around proteins is found in [6]. Despite extensive research on protein dynamics the in- vestigations of elementary conformational motions are rare. Specific molecular mechanisms, including the in- volvement of water molecules, that drive the conforma- tional moves are highly demanded as they ultimately de- fine all rearrangements of proteins as a whole. In this work we analyse molecular dynamics (MD) simulated peptide focusing on the moments of ele- mentary conformational changes including explicit wa- ter molecules. We show that water indeed drives the changes, we elucidate the specific mechanisms of this phe- nomenon. We study a zwitterion L-alanyl-L-alanine, Fig. 1, a very convenient model because i) the conformation of the molecule is completely defined by the two dihedral angles ψ and φ, ii) in water the conformation ψ ≈ 2.5, φ ≈ −2.2 radians is prevalent, however very rare transi- tions to two other metastable conformations take place, and iii) the transitions only happen in water because of the molecule's charged ends. The three well separated metastable states, clearly vis- ible on the density of states, Fig. 1, allow to introduce a simple natural discretisation of the conformational states. By also discretising time with a step ∆t the continuous MD trajectory can be converted into a string of symbols {si}, i = 0 . . . N , where si equals to 'A', 'B', or 'C' de- pending on where the trajectory point falls at the time moment ti, N is the number of time steps. FIG. 1: Left: L-alanyl-L-alanine zwitterion and the nor- malised density of its conformations (Ramachandran plot) formed by a 1µs trajectory; right: the same probabilities em- phasising the presence of two minor conformations and the partitioning for symbolisation An important problem is how to identify the moments of transition. The described discretisation of the confor- mational states defines boundaries of the states. How- ever, they delineate the density of conformations aver- aged over the whole trajectory. As we are interested in the dynamically metastable states, that is the configu- rations in which the trajectory spends significantly more time compared to the time it spends in transitions be- tween the configurations, a dynamical model capable of identifying the transitions is required. Also, the trajec- tory does not go directly from one state to another, in- stead it winds in a complicated manner, often crossing the states boundaries many times. One of the most popular description of protein dynam- ics lately is the Markov State Model (MSM) [7]. The model specifies the probabilities of each of the discrete states as well as the probabilities of the transitions be- tween them. The MSM transition matrix can be cal- culated from the MD trajectory by counting the state changes and for the studied molecule it is A B C A 0.997 0.002 0.001 B 0.261 0.737 0.002 C 0.097 0.001 0.901 where the value at row i and column j gives the proba- bility of going to state j being currently at state i. MSM provides many useful quantities describing the system [7]. In particular, it tells that the transition A → B happens once in 2.8ns on average. We will concentrate on this specific transition for the rest of the paper. The moments of the A → B transitions within the MSM framework are the values of ti for which si is equal to B while being equal to A at the previous time mo- ment ti−1. The time precision of identifying the transi- tions is ∆t, which is insufficient for studying the transi- tions themselves. The model is valid only for relatively large time steps, that follows from the requirement for the transitions to be history independent (statistically uncorrelated). For this peptide the minimal valid time step is ≈ 6ps. This value is of the same order as the period of fluctuations within each conformational state and, most importantly, this is approximately the dura- tion of the process of the trajectory passing from state to state. Therefore, the MSM has to be augmented in order to be able to describe the dynamics at significantly shorter time steps. For this purpose we build a variant of the hidden Markov model using the same configurational states of Specifically, we use the 'ǫ-machine' by the peptide. Crutchfield et al [8, 9]. Instead of the conformational states, si, themselves (A,B,C) we consider the l-long sequences of states ←−s i ≡ {si−l+1 . . . si−2si−1si}. The advantage of such description is that for a small time step, even if the original states are correlated over sev- eral steps, for long enough sequences ←−s i, these new states (the sequences) are uncorrelated. We, therefore, can build a Markov model on these new states. The Markov property on the original states is fulfilled for ∆t > 6ps. For ∆t = 5ps the sequences of at least two time steps are required to build the hidden Markov model. The model is shown in Fig. 2. Here the states '0', '2', and '1' correspond to the conformational states A, B, and C since they mostly consist of the sequences 0 ≡ AA, 2 ≡ BB, and 1 ≡ CC respectively. An additional state '2' describes the transition process from A to B. It consists of the sequence AB and has two main transitions from it. With the probability 0.643 the following symbol is B and the next state is 3 which means that the system is transferred to the conformational state B. There is, however, a significant probability of 0.351 for the next symbol to be A, which describes the return to state 0 or the original conformation A. This analysis illustrates the advantage of the hid- den Markov model: it elucidates the mechanism of the A → B transition. It also explains where the non-Markov property comes from and gives the time scale limit at which different pathways of the conformational transi- tion start to differentiate from each other. Using the hidden Markov model the time step can be reduced to 0.3ps, the model is described in Fig. 3. The 2 FIG. 2: Hidden Markov states for the time step of 5ps; left: conformations corresponding to all time frames belonging to the hidden Markov states; right: the hidden Markov states and transitions between them, the labels on the arcs indicate the symbol following the state as the result of the transition and the transition probability transitions are the moments when the system enters state 9 being before in any other state. The model predicts that the transitions last on average for ≈ 1ps, Fig. 3, right. Thus, the 0.3ps precision in identifying them is satisfactory. This gives us a tool to investigate what happens at different moments before the transition. In particular, we can study the behaviour of water. For this we collect the time frames at specific times before the transitions, Fig. 4. We calculate the density of oxygen (hydrogen) atoms by averaging over the selected time frames. The obtained field ρ(x, t) gives the probability of finding water atoms at various locations x around the peptide at times t in advance of the transition. The transition from conformation A to conformation B corresponds to ≈ 180◦ flip of the N H3 group (the right hand side of the molecule in Fig. 5). Both ends of the 3 FIG. 3: Left: the ǫ-machine for the time step 0.3 ps, the length of the sequences is l = 4, state '0' corresponds to the original state 'A' (mostly consists of the sequences AAAA), similarly, state '9' corresponds to 'B', state '9' can only be reached from state '0' via the states '2', '7', '4', '8' that describe the mecha- nism (pathways) of the A → B transition; right: three typical cases during the transition: direct transition (top, probability 0.61), transition with several recrossings (middle, probability 0.04), failed attempt of transition (bottom, probability 0.07), the probabilities of these cases are given assuming that the probability of going from state '0' to state '9' by any possible route is 1 FIG. 4: Collecting the time frames for the 'time before tran- sition' statistics; the dots on the 'time' axes are the transition moments molecule posses charges, negative on the CO2 and pos- itive on the N H3 sites, which leads to relatively strong attachment of water molecules at the ends. Hydrogen bonded water molecules to the oxygens and to the hy- drogens form the dense areas of water corresponding to more rigid hydrogen bonds network, that is more stable structures. This is an intuitively clear result. More interesting is the moment just before the transition, Fig. 5, middle. Even though the overall structure of the dense areas re- mains similar to the A state, evidently, the total area of the dense water is significantly reduced. We quan- tify this decrease by measuring the volume of the dense areas depicted in Fig. 5. The result as the function of time before the transition is given in Fig. 6. During the time period when the rotation of the N H3 group is most significant, from ≈ 9ps to ≈ 1ps, the size of the dense areas of water is more than 2 times for oxygen and 3 times for hydrogen smaller compared to those during the stable periods of conformation A. This corresponds to more diffuse character of the hydrogen bonded network of water molecules as they tend to appear at different, FIG. 5: Density isosurfaces at 0.7 am (red) and 0.1 am A (top), B (bottom), and ≈ 3ps before the transition A3 (average: 0.5) of oxygen A3 (average: 0.08) of hydrogen (blue) for states FIG. 6: The size of the high density areas (the isosurface values are the same as in Fig. 5 for oxygen (red) and hydrogen (black)) during the transition from state A to state B (the points for state B are artificially shifted to the value of ≈ 0.05ps less concerted locations for different transitions. However, the most surprising effect following from this analysis is that the density of water starts reducing as early as ≈ 50ps in advance of the transition. This is almost 10 times earlier than the actual conformational change of the peptide! The above analysis reveals the changes in water den- sities but it does not show which part of it is directly correlated with the changes in the peptide angles. We analyse the dynamical correlations between the dihedral angles and the water density using the Linear Stochastic Estimation (LSE) model (originally developed for visual- ising coherent structures in turbulent flows [10] and the identification of noise sources in turbulent jets [11]). The density field ρ(x, t) is converted to a time fluctuat- ing field by subtracting the time average ¯ρ(x): ρ′(x, t) = ρ(x, t) − ¯ρ(x). In the LSE framework an approximation ρ(x, t) = α ¯φ(t)+ β ¯ψ(t) to ρ′(x, t) is found that minimises the residual error hρ′ − ρit [12], where ¯φ(t), ¯ψ(t) are the angles averaged over all the frames at time t and α, β are constants. ρ(x, t) represents the density time fluc- tuations correlated with the angles. It is calculated by solving for α, β the system of linear equations, obtained from hρ′(x, t) ¯φ(t)i = h{α ¯φ(t) + β ¯ψ(t)} ¯φ(t)i (and simi- larly for ¯ψ(t)) in the assumption of the zero mean of the uncorrelated part of the density fluctuations: αh ¯φ(t) ¯φ(t)i + βh ¯ψ(t) ¯φ(t)i = hρ′(x, t) ¯φ(t)i αh ¯φ(t) ¯ψ(t)i + βh ¯ψ(t) ¯ψ(t)i = hρ′(x, t) ¯ψ(t)i. The density fluctuations field ρ′(x, t) and its part cor- related with the peptide's conformation ρ(x, t) are shown in Fig. 7 for several representative time moments. The water fluctuations (right column) are significantly stronger just before the transition (2.4ps) compared to the stable period (≈ 100ps). Very surprisingly the con- ditionally averaged water fluctuations (left column) are virtually uncorrelated with the peptide at all times ex- cept for the short period immediately before the tran- sition (some fluctuations are also noticeable as early as ≈ 20ps before the transition, which agrees with the on- set of the density decrease in Fig. 6). Interestingly, at 0ps, when the transition process is complete, the water density becomes uncorrelated with the peptide, similar to the stable periods. However, the fluctuations of it remain strong, only slightly weaker than at 2.4ps. We explain this effect by the large inertia of the water shell, that needs relatively long time for the fluctuations to settle down. The fact that these post-transition water fluctuations are decoupled from the peptide emphasizes the discovered phenomenon of strong water-peptide in- teractions precisely during the transition process. Summarising, we have found that (i) ≈ 5ps before the transition, when the dihedral angles change the most, the water density significantly reduces; (ii) the change of water density begins at ≈ 50ps before the transition, 10 times earlier than the changes in the angles (iii) dur- ing the transition the dynamics of water density becomes 4 highly correlated with the dynamics of the angles; and (iv) these correlations are completely absent during the stable conformation periods. We conclude that water and the peptide behave as an integral dynamical system. During the conformational transition the peptide and the surrounding water un- dergo transitions together. This is in contrast to the metastable periods when their dynamics is essentially de- coupled. The transition is characterised by a more dif- fuse hydrogen bonds network of water. The changes in the peptide are substantially delayed in time. Thus, it is likely that water drives the whole process of conforma- tional transitions. ∗ Electronic address: [email protected] † Electronic address: [email protected] [1] B. Born, H. Weingartner, E. Brndermann, and M. Havenith, JACS 131, 3752 (2009), pMID: 19275262. [2] S. E. Pagnotta, S. Cerveny, A. Alegria, and J. Colmenero, Phys. Chem. Chem. Phys. 12, 10512 (2010). [3] Q. Johnson, U. Doshi, T. Shen, and D. Hamelberg, J. Chem. Theory Comput. 6, 2591 (2010). [4] L. Zhang, Y. Yang, Y.-T. Kao, L. Wang, and D. Zhong, JACS 131, 10677 (2009), pMID: 19586028. [5] H. Frauenfelder, G. Chen, J. Berendzen, P. W. Fenimore, H. Jansson, B. H. McMahon, I. R. Stroe, J. Swenson, and R. D. Young, PNAS 106, 5129 (2009). [6] S. Ebbinghaus, S. J. Kim, M. Heyden, X. Yu, U. Heugen, M. Gruebele, D. M. Leitner, and M. Havenith, PNAS 104, 20749 (2007). [7] C. Schuette, A. Fischer, W. Huisinga, and P. Deuflhard, J. Comput. Phys. 151, 146 (1999), ISSN 0021-9991. [8] J. P. Crutchfield and K. Young, Phys. Rev. Lett. 63, 105 (1989). [9] D. Nerukh, V. Ryabov, and R. C. Glen, Phys. Rev. E 77, 036225 (2008). [10] R. J. Adrian, Appl. Sci. Res. 53, 291 (1994), ISSN 0003- 6994, 10.1007/BF00849106. [11] F. Kerherve, P. Jordan, J. Delville, C. Bogey, and D. Juve, in Proceeding of the 16th AIAA/CEAS Aeroa- coustics Conference, AIAA, Reston, Virginia, USA (2010), pp. AIAA 2010 -- 3965. [12] A. Papoulis and S. Pillai, Probability, Random Variables and Stochastic Processes (McGraw Hill, 2002). 5 FIG. 7: The xy cross-section (the value of the z coordinate is chosen such that the cutting plane passes through the centre of mass of the peptide) of the density fluctuation field ρ′(x, t) (right), and its part correlated with the peptide's conforma- tion ρ(x, t) (left) for oxygen; the time before transition are, from top to bottom, 94.8ps, 21ps, 2.4ps, and 0ps; the density scale is restricted to the −0.2 . . . 0.2 interval for clarity, the maxima of the peaks reach the values of 0.78 and -0.42
1211.1126
1
1211
2012-11-06T07:13:37
Cargo transportation by two species of motor protein
[ "physics.bio-ph", "q-bio.SC" ]
The cargo motion in living cells transported by two species of motor protein with different intrinsic directionality is discussed in this study. Similar to single motor movement, cargo steps forward and backward along microtubule stochastically. Recent experiments found that, cargo transportation by two motor species has a memory, it does not change its direction as frequently as expected, which means that its forward and backward step rates depends on its previous motion trajectory. By assuming cargo has only the least memory, i.e. its step direction depends only on the direction of its last step, two cases of cargo motion are detailed analyzed in this study: {\bf (I)} cargo motion under constant external load; and {\bf (II)} cargo motion in one fixed optical trap. Due to the existence of memory, for the first case, cargo can keep moving in the same direction for a long distance. For the second case, the cargo will oscillate in the trap. The oscillation period decreases and the oscillation amplitude increases with the motor forward step rates, but both of them decrease with the trap stiffness. The most likely location of cargo, where the probability of finding the oscillated cargo is maximum, may be the same as or may be different with the trap center, which depends on the step rates of the two motor species. Meanwhile, if motors are robust, i.e. their forward to backward step rate ratios are high, there may be two such most likely locations, located on the two sides of the trap center respectively. The probability of finding cargo in given location, the probability of cargo in forward/backward motion state, and various mean first passage times of cargo to give location or given state are also analyzed.
physics.bio-ph
physics
1 Cargo transportation by two species of motor protein Yunxin Zhang1,∗, 1 Shanghai Key Laboratory for Contemporary Applied Mathematics, Laboratory of Mathematics for Nonlinear Science, Centre for Computational Systems Biology, School of Mathematical Sciences, Fudan University, Shanghai 200433, China. ∗ E-mail: [email protected] Abstract The cargo motion in living cells transported by two species of motor protein with different intrinsic directionality is discussed in this study. Similar to single motor movement, cargo steps forward and backward along microtubule stochastically. Recent experiments found that, cargo transportation by two motor species has a memory, it does not change its direction as frequently as expected, which means that its forward and backward step rates depends on its previous motion trajectory. By assuming cargo has only the least memory, i.e. its step direction depends only on the direction of its last step, two cases of cargo motion are detailed analyzed in this study: (I) cargo motion under constant external load; and (II) cargo motion in one fixed optical trap. Due to the existence of memory, for the first case, cargo can keep moving in the same direction for a long distance. For the second case, the cargo will oscillate in the trap. The oscillation period decreases and the oscillation amplitude increases with the motor forward step rates, but both of them decrease with the trap stiffness. The most likely location of cargo, where the probability of finding the oscillated cargo is maximum, may be the same as or may be different with the trap center, which depends on the step rates of the two motor species. Meanwhile, if motors are robust, i.e. their forward to backward step rate ratios are high, there may be two such most likely locations, located on the two sides of the trap center respectively. The probability of finding cargo in given location, the probability of cargo in forward/backward motion state, and various mean first passage times of cargo to give location or given state are also analyzed. Introduction Motility is one of the basic properties of living cells, in which cargos, including organelles and vesicles, are usually transported by cooperation of various motor proteins [1, 2], such as the plus-end directed kinesin and minus-directed dynein [3–5]. Experiments found that, using the energy released in ATP hydrolysis [6–9], these motors can move processively along microtubule with step size 8 nm and in hand- over-hand manner [10–12]. Although numerous experimental and theoretical studies have been done to understand this cargo transportation process, so far the mechanism of which is not fully clear. In [13], one basic model is presented by assuming cargo is transported by only one motor species and all the motors share the external load equally. Then in [14], one more realistic tug-of-war model is designed, in which the cargo is assumed to be transported by two motor species with opposite intrinsic directionality, and motors can reverse their motion direction under large external load. According to some experimental phenomena this tug-of-war model seems reasonable [15, 16]. In either of the models given in [13, 14], the only interaction among different motors is that, motors from the same species share load equally and motors from different species act as load to each other. In [17–19], some complicated models are presented, in which interactions among motors are described by linear springs. Recent experiments found that the tug-of-war model might not be reasonable enough to explain some experimental phenomena, so several new models are designed to try to understand the mechanism of cargo motion by multiple motors [20–26]. Finally, more discussion about cargo transportation in cells can be found in [27–35]. In recent experiment [36], by measuring cargo dynamics in optical trap, Leidel et al. found cargo 2 motion along microtubule has memory. Cargo is more likely to resume motion in the same direction rather than the opposite one. This finding implies that, cargo location in the next time depends not only on its present location but also on how it reaches the present location. The behavior of cargo depends on its motion trajectory, which is different from the assumptions in previous models. In this study, one model for cargo motion with memory will be presented. But for simplicity, we assume that the cargo has only a little memory, it can only remember the motion direction in its last step. The description and theoretical analysis of the model with memory will be first given in the next section, and then corresponding results will be presented in the following section. Results will be summarized in the final section. Model for cargo motion with memory In this study, the cargo is assumed to be tightly bound by two motor species: plus-end (or forward) motors and minus-end (or backward) motors. The forward and backward step rates of each plus-end motor are u and w, and the forward and backward step rates of each minus-end motor are f and b. Obviously u ≫ w but b ≫ f when the external load is low, since the intrinsic directionalities of the two motor species are opposite to each other, and the intrinsic motion direction of plus-end motor is plus-end directed (i.e. to the plus-end of microtubule), but the intrinsic motion direction of minus-end motor is minus-end directed (i.e. to the minus-end of microtubule). By assuming that all motors from the same motor species share the load equally, we only need to discuss the simplest cases in which the cargo is transported by only one plus-end motor and one minus-end motor. For example, if there are k plus-end motors, the total external load is Fc, the forward and backward step rates of one single plus-end motor are uc and wc, and the motor step size is lc. Then these k plus-end motors can be effectively replaced by one single plus-end motor with load F = Fc/k, step rates u = kuc and w = kwc, and step size l0 = lc/k. Since the experiments in [36] showed that, the number of motors moving the cargo is usually the same in both directions, this study also assumes the step sizes of the plus-end motor and minus-end motor are the same (note, the step size of single plus-end motor kinesin and step size of single minus-end motor dynein are the same l0 ≈ 8 nm [2, 9, 12]). This study will mainly discuss two special cases: (I) Cargo moves under constant external load. In vitro, this constant load may be applied by one feedback optical trap, or In vivo, this constant load may be from the viscous environment with invariable drag coefficient. (II) Cargo moves in one fixed optical trap, this case is easy to be performed experimentally, and so the corresponding theoretical results are easy to be verified. Cargo Motion under constant load For the sake of convenience, the cargo is said to be in plus-state n+ if it reached its present location n by one forward step from location n − 1. Similarly, the cargo is said to be in minus-state n− if its previous step is minus-end directed, see Fig. 1(a) for the schematic depiction. In plus-state, the forward step rate is higher than backward step rate u > w, but in minus-state the forward step rate is lower than backward step rate f < b. So in plus-state, the cargo is more likely to move forward, but in minus-state, the cargo will be more likely to move backward. For example, for a cargo in location n, if its previous step is plus-end directed, from either plus-state n+ − 1 or minus-state n− − 1 to location n, then in the next step the cargo will be more likely to move to location n + 1 (plus-state n+ + 1), since the cargo is now in plus-state n+ and its forward step rate u is higher than its backward step rate w. On the contrary, if it got to its present location n from location n + 1 (either from plus-state n+ + 1 or from minus-state n− + 1), then in the next step the cargo will be more likely to move to location n − 1 (minus-state n− − 1), since the cargo is now in minus-state n− and its backward step rate b is higher than its forward step rate f . This behavior means that the cargo can remember its motion direction of its last step. Let p, ρ be probabilities of cargo in plus-state and minus-state respectively, then dp/dt = f ρ − wp = −dρ/dt. Using the normalization condition p + ρ = 1, its steady state solution can be obtained as follows p = f /(f + w), ρ = w/(f + w). Let Uef f = up + f ρ, Wef f = wp + bρ, then the mean velocity of cargo can be obtained as follows V = (Uef f − Wef f )l0 = [(u − w)p + (f − b)ρ]l0 = (uf − wb)l0/(f + w), where l0 is the step size of cargo. The probabilities that cargo steps forward and backward are then p+ = Uef f Uef f + Wef f = f (u + w) , p− = 1 − p+ = f (u + w) + w(f + b) w(f + b) . f (u + w) + w(f + b) 3 (1) (2) (3) (4) Finally, the external load F dependence of rate u, w, f, b can be given by the following Bell approxi- mation [37–40], u = u0e−ǫ0F l0/kB T , w = w0e(1−ǫ0)F l0/kB T , f = f0e−ǫ1F l0/kB T , b = b0e(1−ǫ1)F l0/kB T . (5) Where ǫ0 and ǫ1 are load distribution factors for the plus-end motor and minus-end motor, respectively. kB is Boltzmann constant, and T is the absolute temperature. Cargo Motion in one fixed optical trap This special case is schematically depicted in Fig. 1(b). For convenience, the center of optical trap is assumed to be fixed at location 0. For this case, the potential of cargo depends on its location n. The potential difference between location n and location n + 1 is ∆Gn = κ[(n + 1)l0]2/2 − κ(nl0)2/2 = κ(n + 1/2)l2 0. Similar as in [19], at location n, the forward and backward step rates un and wn of cargo in plus-state, as well as the step rates fn and bn of cargo in minus-state, can be obtained as follows, un = ue−ǫ0∆Gn/kB T , wn = we(1−ǫ0)∆Gn−1/kB T , fn = f e−ǫ1∆Gn/kB T , bn = be(1−ǫ1)∆Gn−1/kB T . (6) Where u, w, f, b are cargo step rates when there is no optical trap and any other external load, which satisfy u ≫ w, b ≫ f . For simplicity, this study assumes that ǫ0, ǫ1 are independent of cargo location n. Let pn, ρn be the probabilities of finding cargo in plus-state n+ and minus-state n−, respectively. One can easily show pn, ρn are governed by the following equations dpn/dt = un−1pn−1 + fn−1ρn−1 − (un + wn)pn, dρn/dt = wn+1pn+1 + bn+1ρn+1 − (fn + bn)ρn. (7a) (7b) The steady state solution of Eqs. (7a, 7b) are as follows (for details see Sec. A of the supplemental materials) pn ="n−1 Yk=0(cid:18) (fk + bk)uk pn =" 0 Yk=n+1(cid:18) (uk + wk)bk−1 (uk+1 + wk+1)bk(cid:19)# p0, (fk−1 + bk−1)uk−1(cid:19)# p0, for n ≥ 1, for n ≤ −1, (8a) (8b) ρn = un bn pn = ρn = un bn pn = un bn "n−1 Yk=0(cid:18) (fk + bk)uk bn " 0 Yk=n+1(cid:18) (uk + wk)bk−1 (uk+1 + wk+1)bk(cid:19)# p0, (fk−1 + bk−1)uk−1(cid:19)# p0, un ρ0 = p0. u0 b0 for n ≥ 1, for n ≤ −1, 4 (8c) (8d) (8e) Where p0 can be obtained by the normalization condition P+∞ The probability of finding cargo in plus-state is p =P+∞ in minus-state is ρ =P+∞ +∞ n=−∞ ρn. The mean locations of cargo in plus-state and in minus-state are n=−∞(pn + ρn) = 1. n=−∞ pn, and the probability of finding cargo hn+i = npn/p, hn−i = nρn/ρ, Xn=−∞ +∞ Xn=−∞ respectively. The mean location of cargo is hni = +∞ Xn=−∞ n(pn + ρn) = phn+i + ρhn−i. (9a) (10) Specially, for the symmetric cases u = b, w = f , i.e. the cargo is transported by two motors with the same step rates but different intrinsic directionality, one can verify that ρn = p−n and consequently ρ = p, hn−i = −hn+i, hni = 0. The external load dependence of rates un, wn, fn, bn [see Eq. (6)] means that, for a cargo towed by two motors in one fixed optical trap there are two critical values of the cargo location n, nc+ =(cid:24) kBT κl2 0 ln u w + 1 2 − ǫ0(cid:25) , nc− =(cid:22) kBT κl2 0 ln f b + 1 2 − ǫ1(cid:23) , (11) where ⌈x⌉ is the smallest integer number which is not less than x, ⌊x⌋ is the biggest integer number which is not bigger than x. The step rates of plus-end motor satisfy un > wn for n < nc+, and un ≤ wn for n ≥ nc+. Similarly, the step rates of minus-end motor satisfy bn > fn for n > nc−, and bn ≤ fn for n ≤ nc−. The intrinsic directionality of plus-end motor (u ≫ w) implies nc+ > 0, and the intrinsic directionality of minus-end motor (b ≫ f ) implies nc− < 0. Generally, the critical values nc+ and nc− are different with the mean locations hn+i and hn−i. In the following of this section, various mean first passage time (MFPT) problems about the cargo motion in fixed optical trap will be discussed. Mean first passage time to one of the plus-state Let tl n and τ l n be MFPTs of cargo from plus-state n+ and minus-state n− to plus-state l+ respectively, then tl n and τ l n satisfy [41, 42] wnτ l n−1 − (un + wn)tl n + untl n+1 = −1, for n 6= l, with one boundary condition tl From Eq. (12a) one can easily get bnτ l l = 0. n−1 − (fn + bn)τ l n + fntl n+1 = −1, τ l n−1 = un + wn wn tl n − un wn tl n+1 − 1 wn , for n 6= l. (12a) (12b) (13) Substituting (13) into (12b), one obtains 5 bn(cid:20) un + wn wn tl n − un wn tl n+1 − 1 wn(cid:21) − (fn + bn)(cid:20) un+1 + wn+1 wn+1 tl n+1 − un+1 wn+1 tl n+2 − 1 wn+1(cid:21) + fntl n+1 = −1, i.e. where Bntl n − (Bn + Fn)tl n+1 + Fntl n+2 = Cn, Bn = (un + wn)bn wn , Fn = (fn + bn)un+1 wn+1 , Cn = bn wn − fn + bn wn+1 − 1. Note, Eqs. (14, 15) are established for n 6= l − 1, l. Meanwhile, from Eq. (12b) one can get tl n+1 = fn + bn fn τ l n − bn fn τ l n−1 − 1 fn , and then by substituting Eq. (17) into Eq. (12a) one obtains wnτ l n−1 − (un + wn)(cid:20) fn−1 + bn−1 fn−1 τ l n−1 − bn−1 fn−1 τ l n−2 − 1 fn−1(cid:21) + un(cid:20) fn + bn fn τ l n − bn fn τ l n−1 − (14) (15) (16) (17) 1 fn(cid:21) = −1, (18) i.e. where Bnτ l n−2 − ( Bn + Fn)τ l n−1 + Fnτ l n = Cn, Bn = (un + wn)bn−1 fn−1 , Fn = (fn + bn)un fn , Cn = un fn − un + wn fn−1 − 1. (19) (20) Eqs. (18, 19) are established for n 6= l. The procedure of getting MFPTs tl boundary condition tl Eq. (13). (3) Getting τ l (4) Getting τ l supplemental materials). (5) Getting tl as follows l = 0 (see Sec. B of the supplemental materials). (2) Getting τ l l−1 from the special case of Eq. (12b), i.e. bl−1τ l n, τ l n is as follows. (1) Getting tl n for n ≤ l − 1 by Eq. (15) and n for n ≤ l − 2 by l−1 = −1. (see Sec. C of the n for n ≥ l + 1 by Eq. (17). This procedure can be summarized l−2 − (fl−1 + bl−1)τ l l−1 obtained in (3) n for n ≥ l by Eq. (19) and boundary value τ l Eq. (15) =====⇒ tl l=0 tl n(n ≤ l − 1) Eq. (13) =====⇒ τ l n(n ≤ l − 2) Eq. (12b) ======⇒ n=l−1 τ l l−1 Eq. (19) =====⇒ τ l n(n ≥ l) Eq. (17) =====⇒ tl n(n ≥ l + 1). (21) Mean first passage time to one of the minus-state Let ¯tl n and ¯τ l n be the MFPTs of cargo from plus-state n+ and minus-state n− to minus-state l−, respectively. Similar as the discussion in Sec. , the MFPTs ¯tl n and ¯τ l n satisfy the following equations wn ¯τ l n−1 − (un + wn)¯tl n + un¯tl n+1 = −1, bn¯τ l with one boundary condition ¯τ l n−1 − (fn + bn)¯τ l l = 0. From Eq. (22a) one can easily get n+1 = −1, n + fn¯tl for n 6= l, ¯τ l n−1 = un + wn wn ¯tl n − un wn ¯tl n+1 − 1 wn . (22a) (22b) (23) Substituting (23) into (22b), one obtains 6 bn(cid:20) un + wn wn ¯tl n − un wn ¯tl n+1 − i.e. 1 wn(cid:21) − (fn + bn)(cid:20) un+1 + wn+1 wn+1 ¯tl n+1 − un+1 wn+1 ¯tl n+2 − Bn¯tl n − (Bn + Fn)¯tl n+1 + Fn¯tl n+2 = Cn, 1 wn+1(cid:21) + fn¯tl n+1 = −1, with Bn, Fn, Cn given by Eq. (16). Note, Eqs. (24, 25) are established for n 6= l. Meanwhile, from Eq. (22b) one can get ¯tl n+1 = fn + bn fn ¯τ l n − bn fn ¯τ l n−1 − 1 fn , for n 6= l, and then by substituting Eq. (26) into Eq. (22a) one obtains wn ¯τ l n−1 − (un + wn)(cid:20) fn−1 + bn−1 fn−1 ¯τ l n−1 − bn−1 fn−1 ¯τ l n−2 − 1 fn−1(cid:21) + un(cid:20) fn + bn fn ¯τ l n − bn fn ¯τ l n−1 − (24) (25) (26) 1 fn(cid:21) = −1, (27) i.e. Bn ¯τ l n−2 − ( Bn + Fn)¯τ l n−1 + Fn ¯τ l n = Cn, (28) with Bn, Fn, Cn given by Eq. (20). Eqs. (27, 28) are established for n 6= l, l + 1. n, ¯τ l The procedure of getting MFPTs ¯tl boundary condition ¯τ l Eq. (26). (3) Getting ¯tl (4) Getting ¯tl supplemental materials). (5) Getting ¯τ l as follows l = 0 (see Sec. D of the supplemental materials). (2) Getting ¯tl n for n ≥ l + 1 by Eq. (28) and n for n ≥ l + 2 by l+2 = −1, n for n ≤ l by Eq. (25) with boundary value ¯tl+1 obtained in (3) (see Sec. E of the n for n ≤ l − 1 by Eq. (23). This procedure can be summarized l+1 from the special case of Eq. (22a), i.e. −(ul+1 + wl+1)¯tl n is as follows. (1) Getting ¯τ l l+1 + ul+1¯tl Eq. (28) =====⇒ ¯τ l l =0 ¯τ l n(n ≥ l + 1) Eq. (26) =====⇒ ¯tl n(n ≥ l + 2) Eq. (22a) ======⇒ n=l+1 ¯tl l+1 Eq. (25) =====⇒ ¯tl n(n ≤ l) Eq. (23) =====⇒ ¯τ l n(n ≤ l − 1). (29) Mean first passage time to one given location Let T l s be the MFPT of cargo from state s to location l (either plus-state l+ or minus-state l−), then one can easily show that tl k, τ l k, ¯tl k, ¯τ l k, for s = k+ and k < l, for s = k− and k < l, for s = k+ and k > l, for s = k− and k > l. (30) T l s =  It is to say that if k < l, a cargo located at k will first reach plus-state l+ before reaching minus-state l−. On the contrary, if k > l, it will first reach minus-state l−. Finally, the mean oscillation period T of cargo in fixed optical trap can be approximated as follows see Sec. F of the supplemental materials for its expression. T ≈ τ 0 0 + ¯t0 0, (31) 7 Results For cargo motion under no external load, Monte Carlo simulations show that, if the cargo is trans- ported by two symmetric motors, i.e., the plus-end motor and the minus-end motor have the same step rates, u = b, w = f , the cargo will oscillate [Fig. 2(a)]. While for the asymmetric cases, the cargo has non-zero mean velocity [see Fig. 2(b)]. On the other hand, if the cargo is put into one fixed optical trap, and transported by two symmetric motors, it will oscillate around the trap center with relatively high frequency [Fig. 2(c)]. Meanwhile, if the trapped cargo is transported by two asymmetric motors, it will also oscillate but its oscillation center may be different with the trap center [Fig. 2(d)]. Both Monte Carlo simulations and theoretical calculations show that, for a cargo transported by two symmetric motors and put in one optical trap, its oscillation period T decreases with trap stiffness κ, motor forward step rates u = b, and motor backward step rates w = f [Fig. 3(a-c)]. Its oscillation amplitude increases with the motor forward step rates u = b, but decreases with both the motor backward step rates u = b and the trap stiffness κ, since high backward step rates and high trap stiffness will prohibit the cargo from moving too far from the trap center [Fig. 3(d-f)]. Let p = ∞ ∞ pn, ρ = Xn=−∞ Xn=−∞ ρn, P+ = Xn>0 (pn + ρn), P− = Xn<0 (pn + ρn). (32) Then p is the probability of finding cargo in plus-state, P+ is the probability that cargo location n > 0 (the center of optical trap is assumed to be at location 0). The meanings of ρ and P− are similar. Both Monte Carlo simulations and theoretical calculations show that, for a cargo transported by two symmetric motors, the ratios p/ρ and P+/P− are always one, and they do not change with trap stiffness κ, forward step rates u = b, and backward step rates w = f [Fig. S1]. Our results also show that, for cargo motion in optical trap by two asymmetric motors, its oscillation period T decreases with trap stiffness κ and forward step rate u, but may not change monotonically with backward step rate w [Figs. S2(a), S3(a), S4(a)]. But similar as the symmetric cases, cargo oscillation amplitude of the asymmetric cases decreases with trap stiffness κ and backward step rate w, and increases with the forward step rate u [Figs. S2(d), S3(d), S4(d)]. The results in Figs. S3(d), and S4(d) imply that, the maximal location nmax that cargo might reach toward the plus-end of microtubule depends only on the step rates u, w of the plus-end motor, and similarly the minimal location nmin that cargo might reach towards the minus-end of the microtubule depends only on the step rates b, f of the minus-end motor. From the results given in Figs. S2(b,c), S3(b,c), and S4(b,c) one can also see that, different from the symmetric cases given in Fig. S1, both the ratio p/ρ and ratio P+/P− depend on trap stiffness κ, forward step rate u, and backward step rate w. To show more details about the dependence of cargo oscillation on trap stiffness κ and motor step rates, examples of probabilities pn, ρ, and their summation pn + ρn are plotted in Fig. 4 and Fig. S5. For either symmetric cases or asymmetric cases, the probability profiles are flat for low trap stiffness κ, indicating that the cargo can reach a farther location from the oscillation center (i.e., with large oscillation amplitude)[Fig. S5]. Similar changes can also be found with the increase of motor forward step rates u or f [Fig. 4(a, b, d)]. Meanwhile, with the increase of motor backward step rates w or f , the probability profile will become more sharp [Fig. 4(c)]. For the asymmetric cases, the most likely location of cargo may be different from the trap center [Fig. S5(c)]. One interesting phenomenon displayed in Fig. 4(b, d) is that, for either the symmetric cases or the asymmetric cases, when motor forward step rates u, b are high, the summation of probability pn + ρn may has two local maxima, indicating that cargo motion in the positive location (n > 0) is mainly dominated by the plus motor, while its motion in the negative location (n < 0) is mainly dominated by the minus motor. Let Nmax pn , Nmax ρn , N(pn+ρn)max be the locations at which probabilities pn, ρn and their summation pn + ρn reach their maxima, respectively. The results plotted in Fig. 5(a) show that, for symmetric motion, Nmax ρn = −Nmax pn and their absolute values increase with the forward to backward step rate 8 ratio u/w = b/f . The results in Fig. 5(d) show that, for low step rate ratio u/w = b/f , the total probability pn + ρn has only one maximum which lies at the trap center. However, with increase of these has one symmetric bifurcation, and its absolute value (see Fig. 4) increases with these ratios, N(pn+ρn)max step ratios. For asymmetric case [see Fig. 5(b)], Nmax pn increases with step rate ratio u/w, but Nmax ρn is independent of it. Which means that, similar as the properties of nmax and nmin displayed in Figs. S3 and S4, Nmax pn depends only on step rates of the plus-end motor, and Nmax ρn depends only on step has rates of the minus-end motor. For asymmetric cases, with the increase of rate ratio u/w, N(pn+ρn)max also one bifurcation, see Fig. 5(e). But one of the two values (the negative one) does not change with depends only on properties of the minus-end u/w. Which means that, the negative one of N(pn+ρn)max motor. Similarly, the positive one of N(pn+ρn)max depends only on properties of the plus-end motor. So both the properties of amplitude nmax, nmin and the most likely locations Nmax pn , Nmax ρn , N(pn+ρn)max indicate that, the plus-end directed motion of cargo is mainly determined by the plus-end motor, and the minus-end directed motion is mainly determined by the minus-end motor, which is one of the main differences with other tug-of-war models [14,18,19,21], and this result is consistent with the experimental phenomena [15,16,36]. Finally, the results in Fig. 5(c) show that, the absolute values of Nmax pn , Nmax ρn decrease with trap stiffness κ, and Fig. 5(f) shows N(pn+ρn)max does not change with stiffness κ. So trap stiffness can change the oscillation amplitude and the oscillation period (see Figs. 3, S2, and S5), but will not change the most likely location N(pn+ρn)max of the cargo. Further calculations of probabilities p, ρ show that, for the symmetric cases both pmax = ρmin and (p + ρ)min decrease with step rate ratio u/w = b/f , and increase with trap stiffness κ [see Figs. S6(a,d)]. Since with large rate ratio u/w = b/f and small stiffness κ, the cargo will oscillate with large amplitude. For the asymmetric cases, pmax 6= ρmin, pmax decreases but ρmin increases with the step rate ratio u/w (i.e. with the increase of the directionality of the plus-end motor). Since with large rate ratio u/w, the plus-end motor has high directionality, and so the cargo moves fast in the plus-state, which means that the probability pn will be flat with large u/w. The plots in Fig. S6(c) show that, although the total probability pn + ρn has two maxima, with the change of rate ratio u/w, the most likely location of cargo may change from one side of the trap center to another side. n, τ l n, ¯tl n, ¯τ l m, τ l m, ¯τ l m ≤ τ l m ≥ ¯τ l n ≤ ¯tl n, ¯tl n, and T l Finally, several examples of MFPTs tl S9-S12, and examples of MFPTs T l m ≤ ¯τ l n ≤ tl tl n ≥ ¯tl ¯tl step rate ratios u/w and b/f are large, then tl m ≥ ¯τ l m ≥ τ l tl S11(b,c,d), S12(a). n are plotted in Fig. 6(a,b) and Figs. S7, S8(a,b), n± are plotted in Fig. 6(c,d) and Fig. S8(c,d). If m < n < l, then n, and T l m ≥ τ l n, n− . Moreover, if the trap stiffness κ is high and the motor m− for m < n < l, and m− for l < n < m, see Fig. 6(a,c,d) and Figs. S7(a,b), S8(c,d),S9, S10(a), n− . If l < n < m, then tl m, ¯τ l n+ ≥ T l n+ ≤ T l m+ , T l m− ≤ T l m+ , T l m− ≥ T l m, T l m+ ≥ T l m, T l m+ ≤ T l m ≤ ¯τ l m ≤ τ l m, ¯tl n ≥ tl m, τ l m, ¯tl Concluding Remarks Recent experimental observations by Leidel et al. [36] show that, in living cells cargo moves along microtubule with memory, i.e., its motion direction depends on its previous motion trajectory. In this study, such cargo transportation is theoretically studied by assuming that the cargo has the least memory, i.e. its motion direction depends only on its behavior in its last step. The cargo will be more likely to step forward/backward if it came to its present location by one forward/backward step. Two cases are mainly discussed: (I) cargo moves under constant load, and (II) cargo moves in one fixed optical trap. For each cases, two kinds of motion are addressed: (i) symmetric motion, in which cargo is transported by two species of motor protein which have the same forward/backward step rates but with different intrinsic directionality, (ii) asymmetric motion, in which cargo is transported by two species of motor protein with different forward/backward step rates. For the symmetric motion (i) of case (I), the mean velocity of cargo is zero. But, due to the existence of memory, cargo can move unidirectionally for a large distance before switching its direction. One can easily understand that, for the asymmetric motion (ii) of (I), 9 the directionality of cargo with memory is better than that in the usual tug-of-war model by two different motor species [14, 19, 21]. For the motion in one fixed optical trap, i.e. case (II), cargo will oscillate. For the symmetric motion (i), the oscillation center is the same as the trap center, but for the asymmetric motion (ii) , this oscillation center is generally different from the trap center. Usually the oscillation period decreases with the trap stiffness κ and motor step rates. Meanwhile, the oscillation amplitude decreases with trap stiffness κ and motor backward step rates w, f , but increases with motor forward step rates u, b. The probability pn + ρn of finding cargo at location n may have only one maximum, which is the same as the trap center for symmetric motion (i) but different with the trap center for asymmetric motion (ii). Meanwhile, the probability pn + ρn may also have two maxima. For symmetric motion (i), these two maxima are located symmetrically on the two side of the trap center, and their corresponding values of probability pn + ρn are the same. However, for the asymmetric motion (ii), these two maxima are generally not symmetrically located around the trap center, and their corresponding probabilities may be greatly different. With the change of ratio of motor forward to backward step rates, the maximum with the larger value of probability pn + ρn may transfer from one side of the trap center to another side. This study will be helpful to understand the high directionality of cargo motion in living cells by cooperation of two species of motor protein. Meanwhile, more generalized model can also be employed to discuss this cargo transportation process, in which the cargo is assumed to have long memory, its forward and backward step rates depend on how long it has kept moving in its present direction. Acknowledgments This study was supported by the Natural Science Foundation of China (Grant No. 11271083), Natural Science Foundation of Shanghai (Grant No. 11ZR1403700), and the National Basic Research Program of China (National "973" program, project No. 2011CBA00804). References 1. Bray D (2001) Cell movements: from molecules to motility, 2nd Edn. Garland, New York. 2. Howard J (2001) Mechanics of Motor Proteins and the Cytoskeleton. Sinauer Associates and Sunderland, MA. 3. Block SM, Goldstein LSB, Schnapp BJ (1990) Bead movement by single kinesin molecules studied with optical tweezers. Nature 348: 348-352. 4. Vale RD (2003) The molecular motor toolbox for intracellular transport. Cell 112: 467-480. 5. Mallik R, Carter BC, Lex SA, King SJ, Gross SP (2004) Cytoplasmic dynein functions as a gear in response to load. Nature 427: 649-652. 6. Hua W, Young EC, Fleming ML, Gelles J (1997) Coupling of kinesin steps to ATP hydrolysis. Nature 388: 390-393. 7. Schnitzer MJ, Block SM (1997) Kinesin hydrolyses one ATP per 8-nm step. Nature 388: 386-390. 8. Coy DL, Wagenbach M, Howard J (1999) Kinesin takes one 8-nm step for each ATP that it hydrolyzes. J Biol Chem 274: 3667-3671. 9. Gennerich A, Carter AP, Reck-Peterson SL, Vale RD (2007) Force-induced bidirectional stepping of cytoplasmic dynein. Cell 131: 952-965. 10 10. Asbury CL, Fehr AN, Block SM (2003) Kinesin moves by an asymmetric hand-over-hand mecha- nism. Science 302: 2130-2134. 11. Toba S, Watanabe TM, Yamaguchi-Okimoto L, Toyoshima YY, Higuchi H (2006) Overlapping hand-over-hand mechanism of single molecular motility of cytoplasmic dynein. Proc Natl Acad Sci USA 103: 5741-5745. 12. Guydosh NR, Block SM (2009) Direct observation of the binding state of the kinesin head to the microtubule. Nature 461: 125-128. 13. Klumpp S, Lipowsky R (2005) Cooperative cargo transport by several molecular motors. Proc Natl Acad Sci USA 102: 17284-17289. 14. Muller MJI, Klumpp S, Lipowsky R (2008) Tug-of-war as a cooperative mechanism for bidirectional cargo transport by molecular motors. Proc Natl Acad Sci USA 105: 4609-4614. 15. Gennerich A, Schild D (2006) Finite-particle tracking reveals sub-microscopic size changes of mi- tochondria during transport in mitral cell dendrites. Phys Biol 3:45-53 3: 45-53. 16. Soppina V, Rai AK, Ramaiya AJ, Barak P, Mallik R (2009) Tug-of-war between dissimilar teams of microtubule motors regulates transport and fission of endosomes. Proc Natl Acad Sci USA 106: 19381-19386. 17. Kunwar A, Vershinin M, Xu J, Gross SP (2008) Stepping, strain gating, and an unexpected force- velocity curve for multiple-motor-based transport. Curr Biol 18: 1173-1183. 18. Kunwar A, Mogilner A (2010) Robust transport by multiple motors with nonlinear force-velocity relations and stochastic load sharing. Phys Biol 7: 016012. 19. Zhang Y (2011) Cargo transport by several motors. Phys Rev E 83: 011909. 20. Rogers AR, Driver JW, Constantinou PE, Jamison DK, Diehl MR (2009) Negative interference dominates collective transport of kinesin motors in the absence of load. Phys Chem Chem Phys 11: 4882. 21. Driver J, Rogers A, Jamison D, Das R, Kolomeisky A, et al. (2010) Coupling between motor proteins determines dynamic behaviors of motor protein assemblies. Phys Chem Chem Phys 12: 10398-10405. 22. Driver JW, Jamison DK, Uppulury K, Rogers AR, Kolomeisky A, et al. (2011) Productive coop- eration among processive motors depends inversely on their mechanochemical efficiency. Biophys J 101: 386-395. 23. Jamison DK, Driver JW, Diehl MR (2011) Cooperative responses of multiple kinesins to variable and constant loads. J Biol Chem 287: 3357-3365. 24. Uppulury K, Efremov AK, Driver JW, Jamison DK, Diehl MR, et al. (2012) How the interplay between mechanical and non-mechanical interactions affect multiple kinesin dynamics. J Phys Chem B 116: 8846-8855. 25. Kunwar A, Tripathy SK, Xu J, Mattson M, Sigua R, et al. (2011) Mechanical stochastic tug-of- war models cannot explain bidirectional lipid-droplet transport. Proc Natl Acad Sci USA 108: 18960-18965. 26. Bouzat S, Levi V, Bruno L (2012) Transport properties of melanosomes along microtubules inter- preted by a tug-of-war model with loose mechanical coupling. PLoS ONE 7: e43599. 11 27. Julicher F, Prost J (1995) Cooperative molecular motors. Phys Rev Lett 75: 2618-2621. 28. Badoual M, Julicher F, Prost J (2002) Bidirectional cooperative motion of molecular motors. Proc Natl Acad Sci USA 99: 6696-6701. 29. Adachi K, Oiwa K, Nishizaka T, Furuike S, Noji H, et al. (2007) Coupling of rotation and catalysis in F1-ATPase revealed by single-molecule imaging and manipulation. Cell 130: 309-321. 30. Bieling P, Telley IA, Piehler J, Surrey T (2008) Processive kinesins require loose mechanical cou- pling for efficient collective motility. EMBO Reports 19: 1121-1127. 31. Mallik R, Gross SP (2009) Intracellular transport: How do motors work together? Curr Biol 19: R416-R418. 32. Brouhard GJ (2010) Motor proteins: Kinesins influence each other through load. Curr Biol 20: R448-R450. 33. Welte MA (2010) Bidirectional transport: Matchmaking for motors. Curr Biol 20: R410-R413. 34. Hendricks AG, Perlson E, Ross JL, Schroeder HW, Tokito M, et al. (2010) Motor coordination via a tug-of-war mechanism drives bidirectional vesicle transport. Current Biology 20: 697-702. 35. Schroeder HW, Mitchell C, Shuman H, Holzbaur ELF, Goldman YE (2010) Motor number controls cargo switching at actin-microtubule intersections in vitro. Curr Biol 20: 687-696. 36. Leidel C, Longoria RA, Gutierrez FM, Shubeita GT (2012) Measuring molecular motor forces in vivo: Implications for tug-of-war models of bidirectional transport. Biophys J 103: 492-500. 37. Bell GI (1978) Models for the specific adhesion of cells to cells. Science 200: 618-627. 38. Fisher ME, Kolomeisky AB (2001) Simple mechanochemistry describes the dynamics of kinesin molecules. Proc Natl Acad Sci USA 98: 7748-7753. 39. Zhang Y (2009) A general two-cycle network model of molecular motors. Physica A 383: 3465-3474. 40. Zhang Y (2011) Growth and shortening of microtubules: A two-state model approach. J Biol Chem 286: 39439-39449. 41. Redner S (2001) A Guide to First-Passage Processes. Cambridge University Press. 42. Zhang Y (2011) Periodic one-dimensional hopping model with transitions between nonadjacent states. Phys Rev E 84: 031104. Tables Table I. The values of rates u, w, f, b (in unit s−1) and optical trap stiffness κ (pN/nm) used in the plots of Figs. 2-6. The symbol ∗ means that the corresponding parameter is not used in the plot, and symbol X means this parameter is one variable in the corresponding plot. Other parameters used in the plots are ǫ0 = ǫ1 = 0.5, l0 = 8 nm, and kBT = 4.12 pN·nm. The stiffness κ of the trap used in recent experiment of Leidel el al. is around 0.02 − 0.09 pN/nm [36]. 12 f 2 1 1 1 1 1 w 2 2 1 1 1 1 u 5 Fig. 2(a) 5 Fig. 2(b) 20 Fig. 2(c) 20 Fig. 2(d) 10 Fig. 3(a,d) Fig. 3(b,e) X Fig. 4(c,f) Fig. 4(a) Fig. 4(b) b 5 2 20 5 10 X 100 X X 100 10 10 50 50 1 1 1 1 κ ∗ ∗ 0.004 0.001 X 0.05 0.05 0.05 0.05 20 50 15 Fig. 4(c) Fig. 4(d) 1 Fig. 5(a,d) X 1 Fig. 5(b,e) X 1 Fig. 5(c,f) 1 1 Fig. 6(a) 1 Fig. 6(b) Fig. 6(c) 1 Fig. 6(d) 1 10 5 5 30 10 20 30 0.05 15 1 0.05 1 X 0.05 0.05 1 1 X 0.05 1 0.01 1 0.05 1 1 0.05 50 10 5 5 10 10 13 Figure 1. Schematic depiction of the model discussed in this study to explain the cargo motion with memory. (a) is for cargo motion under constant load, and (b) is for cargo motion in one fixed optical trap. At any location n, the cargo may be in two different states, plus-state n+ and minus-state n−. Cargo in plus-state n+ means it reaches location n from location n − 1, while cargo in minus-state means it is from location n + 1. For a cargo in plus-state n+, its forward and backward step rates are u and w respectively. But for a cargo in minus-state n−, it has different step rates f and b. For the constant load cases (a), u > w and b > f mean that, if the cargo is in plus-state n+ it will be more likely to move forward to location n + 1. Otherwise, it will be more likely to move backward to location n − 1. 14 0 l / x n o i t a c o l o g r a c 0 l / x n o i t a c o l o g r a c (a) 40 20 0 (c) −20 −40 −60 0 40 20 0 −20 −40 0 (b) 60 50 40 30 20 10 0 −10 0 100 (d) 80 60 40 20 0 0 50 100 150 200 250 300 50 100 150 200 250 300 time t (s) 20 40 60 80 100 50 100 150 200 250 300 time t (s) Figure 2. Trajectory samples of cargo motion by two motors under constant load (a, b), and in one fixed optical trap (c, d). For the symmetric cases (where the step rates of the plus motor are the same as the ones of the minus motor, i.e. u = b, w = f ), the cargo will oscillate around its initial location (a). While for the asymmetric cases, the cargo will have nonzero mean velocity (b). If the cargo is put in one fixed optical trap and transported by two symmetric motors, it will oscillate around the trap center (c). But for the asymmetric cases, the oscillation center may be different from the trap center. For parameter values used in the simulations see Tab. I. 15 (c) 20 40 60 (f) (a) 9 8 7 6 5 4 3 2 ) s ( T d o i r e p n o i t a l l i c s o 0 0.02 0.04 0.06 0.08 0.1 (d) 50 0 n i m n , x a m n 5 4 3 2 1 0 0 10 5 0 −5 (b) 20 40 60 (e) 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 10 5 0 −5 −50 10−3 10−2 κ (pN/nm) 10−1 −10 100 101 u = b (s−1) −10 100 102 101 w = f (s−1) Figure 3. In fixed optical trap, the mean oscillation period T of cargo decreases with trap stiffness κ, forward rates u = b, and backward rates w = f (in fact, log T decreases almost linearly with log κ, log u = log b, and log w = log f ). The oscillation amplitude nmax − nmin decreases with stiffness κ and backward rates w = f , but increases with forward rates u = b. Here nmax and nmin are the max and min locations that cargo can reaches. The circles and squares are obtained by Monte Carlo simulations. In (a, b, c), the solid curves are obtained by formulation (31). The solid lines in (d) are obtained by nc+, nc− given in Eq. (11), and the solid lines in (e, f) are obtained by nc+ + 3, nc− − 3, respectively. For parameter values see Tab. I. 16 −5 0 5 10 pn ρn pn + ρn 0.1 (b) 0.08 0.06 0.04 0.02 u=50, w=1, f =1, b=50 0 −10 0.15 (d) 0.1 0.05 u=50, w=1, f =1, b=30 0 cargo position n 0 5 −5 0 cargo position n 5 Figure 4. Samples of probability pn and ρn for finding cargo in plus-state and minus-state. For the symmetric cases probabilities pn and ρn are mirror symmetry to each other (a, b, c). Their sum pn + ρn, the probability of finding cargo at location n, might has one maximum [at the center of optical trap, see (a, c)] or two symmetric maximum [see (b)]. (d) is one sample for the asymmetric cases. For parameter values see Tab. I. 0.2 (a) y t i l i b a b o r p 0.15 0.1 0.05 u=10, w=1, f =1, b=10 0 0.4 (c) −5 0 5 y t i l i b a b o r p 0.3 0.2 0.1 0 −5 u=20, w=15, f =15, b=20 17 3 2 1 0 −1 −2 Nmax pn Nmax ρn Nmax pn Nmax ρn (b) 20 40 60 (c) −3 0 80 0.02 0.04 0.06 0.08 0.1 1 0.5 0 −0.5 (e) 20 40 u/w (f) 60 80 −1 0 0.02 0.04 0.06 κ (pN/nm) 0.08 0.1 4 3 2 1 0 −1 −2 −3 −4 0 (a) 4 3 2 1 0 −1 −2 −3 −4 0 (d) x a m ) n ρ + n p ( N 4 3 2 1 0 −1 −2 −3 Nmax pn Nmax ρn 20 40 −4 0 60 4 3 2 1 0 −1 −2 20 40 u/w = b/f −3 0 60 Figure 5. The location Nmax pn , Nmax ρn , N(pn+ρn)max that probabilities pn, ρn and their summation pn + ρn reach their maximum. With the increase of rate ratio u/w = b/f both Nmax pn and Nmax ρn leave far away from the trap center (a). (b) implies that Nmax pn increases with ratio u/w, but Nmax ρn is independent of it. With the increase of trap stiffness κ, both Nmax pn and Nmax ρn come close the the trap center (c). (d, e) show that, with the increase of rate ratio u/w = b/f or rate ratio u/w only, the number of maximum of probability pn + ρn of finding cargo at location n may change. But (f) implies that N(pn+ρn)max is independent of trap stiffness κ. For parameter values see Tab. I. n n t0 τ 0 ¯t0 ¯τ 0 n n −10 0 10 (a) 20 + n T 0 T 0 n− 18 −20 0 20 (b) 40 + n T 0 T 0 n− 6 5 4 3 2 1 0 −40 1 0.8 0.6 0.4 0.2 −5 5 Cargo position n 0 (c) 10 15 0 −15 −10 −5 5 Cargo position n 0 (d) 10 15 2 1.5 1 0.5 ) s ( T P F M 0 −20 0.7 ) s ( T P F M 0.6 0.5 0.4 0.3 0.2 0.1 0 −15 −10 n, τ 0 n to plus-state 0+, MFPTs ¯t0 from state l to location 0 (c, d). For high trap stiffness κ, t0 Figure 6. Samples of MFPTs t0 MFPT T 0 l MPFTs to plus-state 0+, and symmetric relations hold for MFPTs to minus-state 0−, see (a). But for low trap stiffness, all MFPTs t0 n increases with the distance between n and trap center 0, see (b). Which means that, for different trap stiffness κ, the trajectories of cargo from state n+ or n− to state 0+ or 0− are different. (c, d) are MFPTs for one cargo (transported by two asymmetric motors) from state n+ or n− to location 0 (plus-state 0+ or 0−) and location 1 (plus-state 1+ or 1−). The MFPT T 0 n is obtained by formulation (30). For parameter values see Tab. I. n to minus-state 0− (a, b), and n, τ 0 n, ¯t0 n<0 < τ 0 m<0 < τ 0 l≥0 < t0 k>0 for n, ¯τ 0 n, ¯τ 0
1111.4726
1
1111
2011-11-21T04:05:55
Dissipative electro-elastic network model of protein electrostatics
[ "physics.bio-ph", "physics.chem-ph" ]
We propose a dissipative electro-elastic network model (DENM) to describe the dynamics and statistics of electrostatic fluctuations at active sites of proteins. The model combines the harmonic network of residue beads with overdamped dynamics of the normal modes of the network characterized by two friction coefficients. The electrostatic component is introduced to the model through atomic charges of the protein force field. The overall effect of the electrostatic fluctuations of the network is recorded through the frequency-dependent response functions of the electrostatic potential and electric field at the active site. We also consider the dynamics of displacements of individual residues in the network and the dynamics of distances between pairs of residues. The model is tested against loss spectra of residue displacements and the electrostatic potential and electric field at the heme's iron from all-atom molecular dynamics simulations of three hydrated globular proteins.
physics.bio-ph
physics
Dissipative electro-elastic network model of protein electrostatics Daniel R. Martin, S. Banu Ozkan, and Dmitry V. Matyushov∗ Center for Biological Physics, Arizona State University, PO Box 871504, Tempe, AZ 85287-1504 July 9, 2018 ∗Correspondence: [email protected] Abstract ABSTRACT We propose a dissipative electro-elastic network model (DENM) to describe the dynamics and statistics of electrostatic fluctuations at active sites of proteins. The model com- bines the harmonic network of residue beads with overdamped dynamics of the normal modes of the network characterized by two friction coefficients. The electrostatic component is intro- duced to the model through atomic charges of the protein force field. The overall effect of the electrostatic fluctuations of the network is recorded through the frequency-dependent response functions of the electrostatic potential and electric field at the active site. We also consider the dynamics of displacements of individual residues in the network and the dynamics of distances between pairs of residues. The model is tested against loss spectra of residue displacements and the electrostatic potential and electric field at the heme's iron from all-atom molecular dynamics simulations of three hydrated globular proteins. Key words: protein electrostatics; dissipative dynamics; elastic network; electrostatic re- sponse function; loss spectrum Electro-elasticmodel INTRODUCTION 2 Proteins in solution exist not as static structures but instead as dynamic entities interconverting between conformational sub-states on the time-scale reflecting the activation barriers involved in the transitions (1 -- 3). It is becoming increasingly clear that biomolecules are fluctuating ma- chines, using dynamical equilibria between their conformational states to promote their function (2, 4). The flexibility of the protein structure naturally leads to changes in both the charge distri- bution of the protein itself and in the polarization of the interfacial waters following the protein motions. These dynamically fluctuating polarization fields will affect many properties sensitive to electrostatics, including the rates of enzymatic reactions (5). The dynamic nature of proteins, enhanced by the conformational manifold provided by hydration water (1), requires the shift of the focus from statistical averages and corresponding thermodynamic parameters, such as equilibrium Gibbs energies, to entire statistical distributions of observables and their dynamics expressed through time correlation functions. This requirement is certainly the case for the problem of protein electron transfer, where, according to the standard Marcus picture (6), both the average of the energy gap between the donor and acceptor energy levels of the electron and its fluctuation are required to determine the probability of electron tunneling. This is just one example when the knowledge of the entire fluctuation spectrum of the observables is of significant interest (7). Other applications, such as electrostatic fluctuations reported by broad-band dielectric spectroscopy (8), THz absorption (9), and light scattering techniques (10) require either the entire fluctuation spectrum or a set of cumulants of the corresponding experimental observables. With the general focus on the fluctuation spectrum of proteins, we propose here a model for the calculation of the statistics and dynamics of the electrostatic potential and electric field in proteins. Electrostatic interactions are clearly important for protein stability and function (5). Electrostatic solvation and interactions affect the stability of folded proteins and their aggrega- tion and crystallization (11 -- 13). Electrostatics might also play a significant role in promoting high-temperature flexibility of proteins through the coupling of the charge distribution in the protein to the electrostatic fluctuations produced by the protein-water interface (14). The dynamics of electrostatic fluctuations spans an enormous range of time-scales from sub-picoseconds to sub-microseconds, and possibly longer (8, 15, 16). While spectroscopic techniques are capable of recording the ultra-fast fs-to-ps relaxation (17), the fluctuations of the protein-water interface recorded by Mossbauer spectroscopy and neutron scattering cover much longer time-scales from sub-nanoseconds to sub-microseconds (3). Our present model aims at these long, and perhaps even longer, time-scales by coarse-graining the protein into an elastic network of beads, each representing a protein residue. Elastic network models have consistently shown good performance in characterizing global, large-scale motions of proteins (18 -- 27). While the low-frequency portion of the spectrum of protein motions is mostly determined by the distribution of mass and molecular shape (28 -- 30), improvements are still needed to account for deficiencies of networks when localized protein motions are involved (31 -- 34). Electrostatic interactions, on the other hand, are notoriously long-ranged. The electrostatic potential, slowly changing over a nanometer-size biomolecule, effectively averages out local structural variations and is mostly sensitive to the global distri- bution of charge within the molecule. From this perspective, the combination of even a basic Electro-elasticmodel 3 elastic network with the molecular charge distribution might be sufficient to describe the statis- tics of the potential fluctuations and their slow dynamics. We propose here a model combining the distribution of molecular charge from the standard atomic force-fields with the dynamics and statistics of protein motions derived from an elas- tic network. To test the model, we employ all-atom Molecular Dynamics (MD) simulations of three hydrated heme proteins. Previous analysis of these data has emphasized the impor- tance of nanosecond (ns) relaxation modes in the electrostatic fluctuations of the protein ma- trix and the protein-water interface (14, 35). This time-scale is relevant to biological function since heme proteins are typical components of biology's energy chains transporting electrons on nanosecond-to-microsecond time-scales (7, 36). The nanosecond time-scale fluctuations are also consistent with the dynamics of the global motions of the network of residues. An elastic network model naturally fits the physics of the problem. The present contribution is a first step in developing network models of the electrostatic response of hydrated proteins. Here, we fo- cus on the electrostatics of the protein matrix only, leaving the water component of the overall response to future studies. THEORY AND METHODS Elastic network model A coarse-grained elastic network model (ENM) assigns a node to a collection of atoms reducing the computational burden of an all-atom normal-mode analysis. The typical coarse-graining is done on the level of individual aminoacids (1-bead model (37)). The position of the node is defined by the coordinates of the Ca atom. The springs connecting the nodes represent the bonded and non-bonded interactions within an accepted cutoff distance (18 -- 20) or by a distance-dependent force constant (26, 33, 38, 39). The ENM diagonalizes the Hessian matrix H derived from a simplified Hookean potential suggested by Tirion (18). This potential, Ei j = C(ri j − r0,i j)2/2, describes the elongation ri j = ri − r j between nodes i and j in the network characterized by one universal force constant C and the structural information stored in the equilibrium bead positions r0,i (r0,i j = r0,i − r0, j). is a 3N × 3N matrix representing the protein elastic energy as a quadratic form of The Hessian H i j the Cartesian displacements d ra i = ra i − ra 0,i of individual beads in the network E = (C/2)(cid:229) d r d ra i H i j j . i, j (1) Here, i and j run between 1 and N, a indicate the Cartesian projections, and summation over repeated Greek indices is assumed here and throughout. The Hessian is diagonalized by the unitary matrix U producing 3N eigenvalues l m. This standard linear algebra formalism (18, 19) yields the statistical correlator between the displacements of residues i and j in the network ,b hd ra i d re ji = (b C)−1(cid:229) U mi l −1 m U m j, m,g (2) where b = 1/(kBT ) is the inverse temperature. If the network is characterized by a universal force constant and a cutoff, it yields a bell-shaped density of vibrational states. A refinement of that network by adopting a stronger coupling between a b a b b g a g e Electro-elasticmodel 4 covalently bound neighbors splits the density of states into two maxima, in better agreement with all- atom calculations (24, 40, 41). This approximation is adopted in our present calculations: the spring constant was multiplied by a constant factor e = 100 for covalent neighbors. In addition, a uniform cut-off radius of 15 A was used in all calculations. Overdamped network dynamics The standard mechanical ENM outlined above obviously lacks dissipative dynamics. The equations of motion are harmonic, implying oscillatory time correlation functions. In contrast, most correlation functions of hydrated proteins observed by scattering (8, 10, 42) and relaxation (43) techniques are exponential, corresponding to the overdamped (Debye) dynamics, or stretched-exponential (8). Several approaches can be implemented to incorporate dissipative dynamics into the mechanical network of beads. Langevin equations of motion for the ENM potential, within the general framework of the Lamm- Szabo formalism (44), have been suggested (34, 48). This approach, and some early suggestions (45, 46), still requires parameterization (done by fitting the rotational and translational diffusion coefficients from MD) and, in addition, doubles the size of the matrix to be diagonalized. We have adopted here a more straightforward formalism not requiring additional computational resources. Instead of a harmonically oscillating network, we have assumed for each normal mode qm, diago- nalizing the network's Hessian, an overdamped motion described by the dissipative memory kernel z (t) (47). The equation of motion for such overdamped dynamics is Z t 0 z (t − t′)qm(t′)dt′ + l mqm = F(t), (3) where F(t) is an external force. The distinction between Eq. (3) and the Langevin network (34, 48) is worth emphasizing here. In the latter, uncoupled Langevin dissipative equations are first assigned to each bead of the network. However, it was noted that dissipative equations for the beads are likely to become coupled when the Langevin dynamics of beads are consistently derived by integrating out the fast degrees of freedom in the equations of motion (49). Given that the assignment of uncoupled Langevin equations to the individual beads is phenomenological from the onset, one can introduce similar phenomenology at a different level of the theory. Here, in the spirit of the standard normal-mode analysis, we introduce uncoupled dissipative equations for the normal modes (Eq. (3)). This is still a phenomenological assumption requiring further testing, but it reduces to the standard normal-mode analysis in the static limit. When Laplace-Fourier transform (47) is applied to Eq. (3), the displacement response function for collective mode qm follows c m(w ) = hiw z (w ) + l mi−1 (4) where z (w ) is the Laplace-Fourier transform of the friction kernel z (t). Extending this procedure to all normal modes of the network, one obtains the response function of bead displacements , i j (w ) = C−1(cid:229) m mi [l m + iw U (w )]−1U m j. (5) In this equation, the response function χi j(w ), which is a rank-2 tensor, represents the displacement of residue i of the protein due to an oscillatory force with frequency w applied to residue j and propagated through the network to residue i. This response function is the basis of the dissipative electro-elastic network model (DENM) proposed here. At w = 0, Eq. (5) returns the standard result of the ENM given by Eq. (2). c a b g a z g b Electro-elasticmodel 5 Figure 1: Cartoon illustrating the application of the dissipative electro-elastic network model (DENM) to the calculation of the electro-elastic response function at an atomic position in the active site of the protein. The elastic displacements of residues i and j create fluctuations of the electric fields by atomic charges qik and q jk. Electric fields Eik and E jk are produced by the corresponding charges qik and q jk at the active site. The overall electric field of residue i at the active site, E0i, is obtained by summation of all Eik produced by the charges qik of the residue. Electrostatic potential response function The ENM has shown good performance for low-frequency structural/mechanical deformations of indi- vidual proteins and biomolecular assemblies (27). Our goal here is to supplement these low-frequency motions with atomic partial charges to model the dynamics and statistics of electrostatic fluctuations at active sites of proteins. Coarse-grained electrostatics of proteins has been addressed in a number of recent papers (50 -- 52), mostly dealing with long-range protein assembly and interactions (53). These ap- proaches also involve coarse-graining of the charge distribution (51, 52) or generating effective charges to reproduce the protein's external field (54). Given that we are interested in the electrostatic fluctua- tions, which are more sensitive to a local charge distribution, atomic charges from force-field potentials are more consistent with our purpose. Our approach starts with solving the standard ENM problem to obtain a set of displacements d rim for each normal mode m with the eigenvalue l m. Each of these displacements is then applied to all charges qik of residue i, where index k runs over all atoms of residue i (Fig. 1). Displacements d rim in ik = qikd rim which interact with charges of the protein active site. The turn produce dipole moments d µ overall electrostatic effect of the fluctuating protein medium is given as a sum over all contributions from each partial charge of all the residues except for the active site. This cumulative effect of all protein charges is given by the frequency-dependent response function of the electrostatic potential c f (w ) (m) f (w ) = −(cid:229) i, j Ea 0 j i j (w )E 0i. (6) Here, Ea an atom in the active site (Fig. 1) 0i is the a Cartesian projection of the electrostatic field produced by all charges qik of residue i at 0i = −(cid:229) Ea qik(rik − r0) rik − r03 . k (7) Here, r0 is the position of the atom in the active site and rik is the position of charge qik of residue i. c c a b b Electro-elasticmodel 6 Electrostatic field response function The potential response function c f (w ) in Eq. (6) describes an oscillating electrostatic potential pro- duced by the protein matrix in response to placing an oscillatory charge at the position in the active site where the potential is recorded (Fe ion in our MD simulations). Similarly, the response function of the electric field considers the field produced by the protein in response to an oscillating dipole moment m0(t) = m0(w )exp[iw t] placed in the active site. The deformation of the protein matrix induced by this probe dipole results in the electric field E0(w ) at the same site. It can be found by summation over the contributions from all individual dipoles d µik(w ) arising from residues' atomic charges E0(w ) = (cid:229) Tik · d µik(w ). ik (8) Here, Tik is the dipolar tensor connecting the position of the active site with the charge qik of residue i. The dipole moment d µik(w ) at the position of qik is caused by the residue displacement d rik(w ) and can be expressed in terms of the displacement response function d µik(w ) = qik(cid:229) χi j · d F j(w ), j (9) where d F j(w ) = (cid:229) k q jkT jk · m0(w ) is the force caused by the dipole m0(w ) at residue j. We therefore obtain a linear relation between the dipole m0(w ) placed at the active site and the electric field E0(w ) produced by the protein in response to this perturbation. The proportionality coefficient between the external perturbation and the response is the response function given by the rank-2 tensor E (w ) = (cid:229) i, j,k,l qikT ik i j (w )T jl q jl. (10) Here, as above, the summation is done over the repeated Greek indices denoting the Cartesian compo- nents of the corresponding tensors. We will be mostly interested in the trace of the tensor c E(w ) = c E (w ). (11) Response and correlation functions The response functions calculated by the DENM formalism need to be related to time correlation func- tions supplied by the MD trajectories. The dynamics of a general dynamical variable d X (t) = X (t) − hX i is characterized by the time self-correlation function (12) The normalized correlation function SX (t) = CX (t)/h(d X )2i is related to the response function by the fluctuation-dissipation equation which forms the basis of our analysis (55) CX (t) = hd X (t)d X (0)i. c X (z) = b h(d X )2i(cid:2)1 + iz SX (z)(cid:3) , (13) where SX (z) is the Laplace-Fourier transform of S(t) defined in the upper half of the complex plane of z. Since our main focus is on variances of physical properties, we will define the generalized compli- ance for the variable X as follows l X = b h(d X )2i/2 = Z 0 ′′ X (w )(dw /p ). (14) c a b a g c g d d b a a ¥ c w Electro-elasticmodel 7 The same property can be calculated from the w = 0 value of the real part of the response function l X = (1/2)c (15) In case of residue displacements considered as variable X (X = r), l r (cid:181) C−1 is proportional to the inverse force constant of the network springs. As for the compliance of a macroscopic body, defined as the inverse of stiffness, l r depends on the shape and boundary conditions, in contrast to elastic moduli representing material properties. ′ X (0). For the elastic network, the generalized compliance reports on how the softness of the network affects reporting on the potential f at the position of the variance of the variable of interest. In the case of X = f the heme's iron, l = e2l is the reorganization energy of a half redox reaction corresponding to changing the oxidation state of the heme (6, 56). The standard definition of half-reaction reorganization energy involves, instead of one atom, the distribution of the electronic density of the transferred electron over a few atoms of the active site. We simplify this problem here by assuming all electron charge e localized at the centroid of this charge density, the iron atom of the heme. The electrostatic potential at a single atom is a well defined physical property, but its variance is only an approximate representation of the observable reorganization energy of changing the redox state (6). We want to gain insight into the dissipative dynamics of the elastic network and its comparison with X (w ) provides direct access to both the dynamics calculated from MD simulations. The loss function c the set of characteristic relaxation frequencies and their relative contributions. However, it does not weigh the low and high frequencies as they contribute to the variance in Eq. (14). Therefore, in addition to the loss function, we will consider the function ′′ This function is normalized, R w → 0 limit 0 where a X (w ) = 2 X (w ) ′′ ′ X (0) . (16) a X (w )dw = 1, and tends to the characteristic relaxation time ht i in the a X (w ) = (2/p )ht i, limw →0 ht i = Z 0 SX (t)dt. (17) (18) Proteins used in case studies and the simulation protocol Three heme proteins, oxidized form of myoglobin (metmyoglobin, metMB, 1YMB) reduced state of cytochrome B562 (cytB, 256B), and reduced state of bovine heart cytochrome c (cytC, 2B4Z) were simulated by all-atom MD. The parameters of the proteins were taken from CHARMM27 force field (57), and NAMD (58) was used for the trajectories production. Each protein was placed at the center of a cubic box with the side length of ≃ 108 A and solvated with TIP3P waters (59). The number of waters used in simulations were: 32891 (metMB), 33268 (cytB), and 33189 (cytC). The VMD (60) plugin Autoionize was used to neutralize the simulation cell by adding Na+ and Cl− ions to bring the ionic strength of the solution to 0.1. Particle-mesh Ewald with the grid resolution < 1 A was used for electrostatic interactions, and all other non-bonded interactions were calculated within 12 A cutoff. Following energy minimization, each protein was simulated for 5 ns in a NPT ensemble at P = 1 atm and T = 300 K. Temperature and pressure were controlled by using the Langevin dynamics with the damping coefficient of 5 ps−1. The production NVE trajectory was started at the end of each 5 ns trajectory. The NVE ensemble is required for the proper sampling of the long-time dynamics since we found that f p w c c ¥ ¥ Electro-elasticmodel 4 3 MD DENM 2 Å / ae 2 2 i r d 1 0 0 50 100 residue No. 100 . o N e u d i s e r 50 0 0 50 100 residue No. 2 Å / ae 2 j i r d 6 4 2 0 100 . o N e u d i s e r 50 0 0 50 100 residue No. 8 2 Å / ae 2 j i r d 3 2 1 0 Figure 2: Residue mean-square displacements (msd's) h(d ri)2i (a) and variance-covariance matrix hd ri · d r ji of residues' Ca carbons calculated from MD (b) and from DENM (c). The diagonals in (b) and (c) represents the residue msd's h(d ri)2i. The network force constant in the DENM calculations is kBT /C = 1 A2. The DENM values indicated by red squares in (a) need to be multiplied by about a factor of two in order to obtain the best-fit agreement with the MD data. They are separated for a better visibility in the plot. using NPT and NVT ensembles artificially accelerates the observables relaxing on sub-nanosecond to nanosecond time-scale (61). The integration step was 2 fs and simulation frames for the analysis were saved each 0.05 ps. The simulation trajectories were 65 ns (metMB), 100 ns (cytC), and 123 ns (cytB). The force field standard parameters of the heme in the reduced state are taken from CHARMM27 (57). The charge of iron is qFe = 0.24 in the reduced state. No standard parameters are available for the oxidized heme required for the simulations of metmyoglobin. These force field parameters were taken from Autenrieth et al (62). The atomic charge of oxidized Fe is qFe = 1.34 in this parametrization. The iron atom is five-coordinate in its oxidized state. In order to allow the iron to shift out of the porphyrin plane, the harmonic force constant of the bond between Fe and Ne 2 of histidine 93 was adjusted to 65 kcal/mol as suggested in Ref. (63) and also implemented in Ref. (64). RESULTS AND DISCUSSION Statistics and dynamics of residue displacements Consistent with many previous studies, we have found that elastic networks are capable of re- producing the basic pattern of the distribution of mean-square displacements (msd's) along the protein backbone. The left panel in Fig. 2 compares residue msd's from MD with DENM calcu- lations, while two other panels show the maps of variance-covariance matrices. Overall, there is a good agreement between the alteration in residue displacements along the backbone, and corresponding cross-correlations, calculated from the two sets of data. However, the network force constant required for the best fit of the msd from MD simulations, ≃ 0.3 kcal/(mol A2), is somewhat lower than the value of ≃ 1 kcal/(mol A2) typically adopted from fitting the B- factors from crystallography (18 -- 20). This value is also about a factor of two lower than 0.6 kcal/(mol A2) adopted below to globally fit the variances of the field and electrostatic potential fluctuations from MD data for all three proteins studied here. The fitting of the force constant to the absolute msd's from MD makes the network too soft. This outcome is expected since the elastic network cuts high-frequency vibrations from the density of states and thus underestimates the absolute magnitudes of the residue displacements (27). It appears more consistent to use for the purpose of fitting only the portion of the msd AE AE AE Electro-elasticmodel 9 ) (w i 0.5 0.4 0.3 0.2 0.1 0 (a) MD l = 5 ns l = 30 ns ) (w j i 0.01 100 1 /ns-1 0.5 0.4 0.3 0.2 0.1 0 (b) MD l = 5 ns l = 30 ns 0.01 100 1 /ns-1 ′′ i (w )/c Figure 3: Loss spectra c ′ i(0) of individual residues (i = 21, a) and of pairs of residues i j(w )/c ′ ′′ i j(0) (i = 21 and j = 84, b). The results have been obtained from MD (solid lines) and DENM (dashed lines). The calculations are done for cytB with the residues used in the calculations shown in Fig. 4. The DENM calculations are done with z l = 30 ns used for the electrostatic calculations in Fig. 5 (red dashed line) and with z l, a = 0.35, e = 100. l = 5 ns (blue dash-dotted line). The rest of DENM parameters are: z h = 0.006z related to protein's global motions. This component can in fact be separated from the vibrational part in the temperature dependence of the msd since global motions of the protein become observable at high temperatures, above the temperature of the protein dynamical transition (65). This high-temperature portion of the protein msd is about a half of the total magnitude (66), which is close to the factor required to reconcile the force constants from electrostatic variances and msd's. Figure 3 shows the dynamics of positions ri(t) of individual residues within the network and the dynamics of distances between residues ri j(t) = ri(t) − r j(t). It reports the loss functions i (w ) and c ′′ i j(w ) calculated from the self-correlation functions of individual residues ′′ Ci(t) = hd ri(t) · d ri(0)i and self-correlation functions of distances between the residues Ci j(t) = hd ri j(t) · ri j(0)i. (19) (20) These latter correlation functions are experimentally available from FRET (67) recording the evolution of the distance between two residues tagged with chromophores. The comparison of c i (w ) from DENM and MD is illustrated in Fig. 3a for only one residue, i = 21, belonging to cytB protein. The pair of residues with i = 21 and j = 84 is used to calculate Ci j(t) in Fig. 3b. The position of the pair in the backbone of cytB is shown in Fig. 4. Examples of calculations for several additional pairs of residues from cytB can be found in Supporting Materials. ′′ Both correlation functions, Ci(t) and Ci j(t), report the dynamics on roughly two time-scales, seen as two peaks in the their loss functions in Fig. 3. The two peaks represent, with varying weights, the faster local dynamics of individual residues and the slower global dynamics of the network. The elastic network with single-exponential, Debye dynamics does not capture two time- scales of the dynamics from MD simulations. When the Debye memory kernel z (w ) = z 0 is used in the response function c m(w ) in Eqs. (4) and (5), the loss function shows only one relaxation peak, slightly deformed from a simple Debye form by the distribution of network's w c † z z w c † z z c c Electro-elasticmodel 10 Figure 4: Cartoon of cytB showing the residues (red) used to produce the correlation functions Ci(t) i j(w ) (i = 21) and Ci j(t) (i, j = 21,84, Eqs. (19) and (20)) and corresponding loss spectra c shown in Fig. 3. The residues used in producing c i j(w ) are connected by the red solid line. The red spheres representing the two chosen residues are centered at their corresponding Ca carbons. The Fe atom of the heme is rendered as a yellow sphere. i (w ) and c ′′ ′′ ′′ eigenvalues l m. A single friction coefficient evidently misses the fact that energy dissipation decreases with increasing frequency of vibrations. Since the protein network is characterized by two spring constants, lower for non-covalent neighbors within the cutoff and higher for covalent neighbors, it must possess at least two characteristic friction coefficients. Friction kernels typically used in applications are phenomenological (47) and we as well cannot offer a consistent theoretical formalism capturing the complex dynamics of residue dis- placements. We have therefore adopted a phenomenological response function best describing the MD results. Instead of searching for a functional form of z (w ) reproducing MD simulations, we have resorted to a two-Debye form for the entire response function of bead displacements in Eq. (4) c m(w ) = iw a h + l m + iw 1 − a l + l m . (21) Here, the amplitude a represents the relative weight of the high-frequency dissipation with the high-frequency friction z h. Correspondingly, z l is the low-frequency friction, which is larger in magnitude. It is obvious that Eq. (21) satisfies the static limit c m(0) = l −1 m . The two-Debye form of c m(w ) is directly applied to calculate the response function of z z 11 Electro-elasticmodel residue displacements i j(w ) = c i j (w ), (22) where the function c i j (w ) is given by Eq. (5). The loss spectra of residue displacements show two peaks and qualitatively agree with the simulations. The positions and relative heights of the peaks can be adjusted by choosing z h,l and the amplitude a in Eq. (21), as is illustrated in Fig. 3a by dashed and dash-dotted lines. The model can therefore be well parameterized to reproduce the dynamics of displacements of a small subset of residues. However, it fails to discriminate between the dynamics of residues across the protein backbone. The loss functions calculated from MD for a number of residues (see Supporting Materials) display similar two-peaks pattern, but the weights and positions of their peaks vary among the residues. In contrast, the dynamics of displacements of individual residues in the elastic network are mostly driven by the global motions of the entire network. As a result, there is little difference between the loss functions of individual residues calculated with DENM. The same statement applies to the dynamics of distances between residues shown in Fig. 3b. While one can reproduce the loss function of a chosen pair of residues by adjusting the parameters of c m(w ) in Eq. (21), these parameters do not translate to all pairs in the network and instead can be only viewed as an average representation of the pairs dynamics. X (w ), is less significant X (w ) in Eq. (16). The function a X (w ) is in the function a X (w ) due to the 1/w more relevant to problems related to the generalized compliance as defined by Eq. (14). The correct representation of the fast dynamics is therefore of lesser importance for these type of problems. The 1/w X (w ) and a X (w ) explains the relative success of the network models in reproducing the pattern of residue msd's (Fig. 2). Those are obtained r(w ) and are less sensitive to the details of local dynamics of by frequency integration of a individual beads in the network represented by high-frequency peak of their loss functions. The fast component of the dynamics, prominent in the loss function c scaling difference between c scaling of c ′′ ′′ ′′ Dynamics of the electrostatic field and potential Essentially all time correlation functions obtained here from MD simulations show a pattern that is roughly represented by a two-exponential decay, requiring in some cases a third decaying ex- ponent for a better mathematical fit. The relative weights of the fast and slow components of the correlation functions differ depending on the observable. The loss function of the electric field E(w ) at the position of protein's heme is mostly single-exponential, with the low-frequency ′′ peak much exceeding the high-frequency one. This pattern is less uniform for the loss function of the electrostatic potential showing different weights of the low- and high-frequency peaks in c ′ X (0) calculated from the DENM and MD for cytB and metMB proteins. Similarly to the case with the residue displacements, the low-frequency coefficient needs adjustment to reproduce the position of the main loss peak: the value of z l = 10 ns was taken for metMB. Still, the set of parameters taken to reproduce the loss spectrum of electrostatic poten- tial does not perform as well when applied to the electric field (cf. Figs. 5c and 5d). ′′f (w ) among the proteins (35). Figure 5 shows the loss spectra c l = 30 ns was adopted for cytB and z X (w )/c ′′ The relative contribution of the fast dynamics component is reduced in the spectral function a X (w ) describing the frequency variation of the generalized compliance (Eqs. (14 and (16)). c a a a b c Electro-elasticmodel (a) cytB 0.4 0.3 0.2 0.1 ) ( X 0 0.01 0.1 (c) 0.3 100 1 10 , ns-1 metMB ) ( X 0.2 0.1 0.1 1 10 , ns-1 100 12 DENM MD 100 1000 1 10 w, ns-1 DENM MD (b) 0.4 0.2 0 0.01 0.1 (d) 0.4 0.2 0 0.01 0.1 1 10 w, ns-1 ′f (0) (a,c) and c ′′f (w )/c 100 1000 Figure 5: Loss spectra c ′ E (0) (b,d) for cytB (a,b) and metMB (c,d). The results are from MD trajectories (dashed lines) and from the DENM calculations (solid lines). The DENM calculations were done with the two-Debye c m(w ) in Eq. (21). The two-Debye relaxation pa- rameters are: z l , a = 0.35 for metMB. The elastic network is defined with kBT /C = 1 A2, e = 100, and the cutoff radius of 15 A. l, a = 0.35 for cytB and z l = 10 ns, z h = 0.0002z l = 30 ns, z h = 0.006z ′′ E (w )/c cytB E cytC E metMB E 10 1 0.1 0.01 0.01 0.1 1 , ns-1 10 0.01 0.1 1 w, ns-1 10 0.01 0.1 10 100 1 w, ns-1 Figure 6: Functions a X (w ) (ns) for cytB, cytC, and metMB from MD (dashed lines) and DENM (solid f (w ), blue) and electric filed (a E (w ), red) are shown. lines). The results for the electrostatic potential (a A single set of network parameters is used for all three heme proteins: z l, a = 0.35, e = 100. l = 30 ns, z h = 0.006z Figure 6 compares a X (w ) from DENM with MD data for all three globular proteins studied here. In order to show the ability of DENM to describe the entire set of MD data, a single set of network and friction parameters has been assigned to all three proteins. As expected, the performance of DENM is better for a X (w ). The average relaxation time ht i, given by the w → 0 limit of a X (w ) (Eq. (17)), is well reproduced by the network calculations. The deviations for some observables correspond to the cases when the dynamics are strongly dominated by its high-frequency component. For these cases, the fit can be improved by adjusting the weight a of the fast component in Eq. (21) to a higher value. The loss functions c X (w ) shown in Figs. 3, 5, and 6 are normalized with the corresponding values of c ′ X (0) and therefore are not affected by the magnitude of the spring constant C de- scribing the non-covalent interactions within the cutoff distance rc = 15 A. The absolute values of the variances are, however, proportional to kBT /C, as for instance in Eq. (2). We adopted in our calculations the value of kBT /C = 1 A2, which is equivalent to C ≃ 0.6 kcal/(mol A2). The w c † w w c † w w f f f Electro-elasticmodel 13 Table 1: Generalized compliances (Eq. (14)) for the electrostatic potential fluctuations l (also known as reorganization energies of redox half reaction l = e2l f , kcal/mol) and for the electric force, F = eE (l F, kcal/(mol A2)) for the three proteins used in the case study; the network is assigned the force constant of C = 0.6 kcal/(mol A2) for non-covalent springs and e C, e = 100 for covalently bound neighbors. The cutoff radius is 15 A. e2l Protein MetMB CytB CytC (DENM) 85 115 141 e2l (MD) l F (DENM) l F (MD) 228 136 159 436 9 18 243 17 83 results of calculating the generalized compliances l X for electrostatic observables (Eqs. (14) and (15)) with this choice of the force constant are summarized in Table 1. The agreement is semi-quantitative at best. While the value of the force constant can clearly be adjusted in each particular case, this change does not propagate into equally good agreement between l f and l F, where F = eE is the electrostatic force acting on the unit charge placed at Fe of the heme. The requirement to reproduce electrostatic forces might be more stringent than to describe the elec- trostatic potential because of a shorter range and higher sensitivity to the local structure of the former. Since the statistical variances listed in Table 1 are not affected by the modeling of dissi- pative dynamics, the properties of the network itself need to be further fine-tuned. The inclusion of softening of the residue displacements by electrostatic water fluctuations (14) appears to be the first avenue for the model improvement. This addition will allow one to accommodate for the effect of the interface, in addition to the motions dictated by the shape and mass distribution which elastic models capture in the first place (28 -- 30). Incorporating the water response will effectively produce a more heterogeneous distribution of the network force constants. CONCLUSIONS Network models of biomolecules are coarse-grained representations designed to describe their large-scale collective conformational motions (27). They capture the basic topology and pack- ing of residues in a biopolymer and robustly reproduce conformational global motions driven by the distribution of mass and shape (28 -- 30). Electrostatic properties is yet another area where coarse-graining of biomolecules might be efficient. The long range of Coulomb interactions ef- fectively averages out details of the local structure suggesting that one can potentially describe the statistics and long-time dynamics of electrostatic fluctuations by global motions of charges assigned to the elastic network. The demonstration that this approach is in principle consistent with all-atom MD simulations is the main result of this study. Two components are critical to our approach: force-field atomic charges distributed at the residues of the network and two- exponential, overdamped dynamics assigned to each normal mode diagonalizing the network Hessian. Harmonic mechanical motions of the network of beads clearly do not incorporate dissipative dynamics of biomolecules in solution (34, 45, 46) and much is still needed to be done to achieve a physically consistent description of the protein dynamics. The elastic network employed here f f f Electro-elasticmodel 14 assigns weaker springs between all non-covalent neighbors within a cutoff radius and stronger springs between covalent neighbors. Correspondingly, two global friction coefficients are as- signed to each normal mode diagonalizing the network Hessian. This phenomenological model qualitatively captures the two-peak loss spectrum of residue displacements and qualitatively similar loss spectra of the electrostatic potential and electric field. However, the characteristic adjustable friction coefficients used for the electrostatic fluctuations are not transferrable to the network displacements. A new set of parameters is needed when each property is considered separately. This research was supported by the National Science Foundation (DVM, CHE-0910905). CPU time was provided by the National Science Foundation through TeraGrid resources (TG-MCB080116N). References 1. Fenimore, P. W., H. Frauenfelder, B. H. McMahon, and R. D. Young, 2004. Bulk-solvent and b -fluctuations in glasses, control protein and hydration-shell fluctuations, similar to a motion and functions. Proc. Natl. Acad. Sci. 101:14408 -- 14413. 2. Henzlel-Wildman, K., and D. Kern, 2007. Dynamic personalities of proteins. Nature 450:964. 3. Frauenfelder, H., G. Chen, J. Berendzen, P. W. Fenimore, H. Jansson, B. H. McMahon, I. R. Stroe, J. Swenson, and R. D. Young, 2009. A unified model of protein dynamics. Proc. Natl. Acad. Sci. 106:5129 -- 5134. 4. Marlow, M. S., J. Dogan, K. K. Frederick, K. G. Valentine, and A. J. Wand, 2010. The role of conformational entropy in molecular recognition by calmodulin. Nat Chem Biol 6:352 -- 358. http://dx.doi.org/10.1038/nchembio.347. 5. Warshel, A., P. K. Sharma, M. Kato, and W. W. Parson, 2006. Modeling electrostatic effects in proteins. Biochim. Biophys. Acta 1764:1647 -- 1676. 6. Marcus, R. A., and N. Sutin, 1985. Electron transfer in chemistry and biology. Biochim. Biophys. Acta 811:265 -- 322. 7. LeBard, D. N., and D. V. Matyushov, 2010. Protein-water electrostatics and principles of bioenergetics. Phys. Chem. Chem. Phys. 12:15335. 8. Khodadadi, S., J. H. Roh, A. Kisliuk, E. Mamontov, M. Tyagi, S. A. Woodson, R. M. Briber, and A. P. Sokolov, 2010. Dynamics of Biological Macromolecules: Not a Simple Slaving by Hydration Water. Biophys. J. 98:1321 -- 1326. 9. Ebbinghaus, S., S. J. Kim, M. Heyden, X. Yu, U. Heugen, M. Gruebele, D. M. Leitner, and M. Havenith, 2007. An extended dynamical hydration shell around proteins. Proc. Natl. Acad. Sci. 104:20749 -- 20752. Electro-elasticmodel 15 10. Perticaroli, S., L. Comez, M. Paolantoni, P. Sassi, L. Lupi, D. Fioretto, A. Paciaroni, and A. Morresi, 2010. Broadband depolarized light scattering study of diluted protein aqueous solutions. J. Phys. Chem. B 114:8262. 11. Richardson, J. S., and D. C. Richardson, 2002. Natural b -sheet proteins use negative design to avoid edge-to-edge aggregation. Proc. Natl. Acad. Sci. 99:2754. 12. Lawrence, M. S., K. J. Phillips, and D. R. Liu, 2007. Supercharging Proteins Can Impart Unusual Resilience. J. Am. Chem. Soc. 129:10110. 13. Pace, C. N., G. R. Grimsley, and J. M. Scholtz, 2009. Protein ionizable groups: pK values and their contribution to protein stability and solubility. J. Biol. Chem. 284:13285. 14. Matyushov, D. V., and A. Y. Morozov, 2011. Electrostatic fluctuations promote dynamical transition in proteins. Phys. Rev. E 84:011908. 15. Andreatta, D., J. L. P´erez, S. A. Kovalenko, N. P. Ernsting, C. J. Murphy, R. S. Coleman, and M. A. Berg, 2005. Power-law solvation dynamics in DNA over six decades in time. J. Am. Chem. Soc. 127:7270. 16. Tripathy, J., and W. F. Beck, 2010. Nanosecond-Regime Correlation Time Scales for Equilibrium Protein Structural Fluctuations of Metal-Free Cytochrome c from Picosecond Time-Resolved Fluorescence Spectroscopy and the Dynamic Stokes Shift. J. Phys. Chem. B 114:15958 -- 15968. 17. Zhang, L., L. Wang, Y.-T. Kao, W. Qiu, Y. Yang, O. Okobiah, and D. Zhong, 2007. Mapping hydration dynamics around a protein surface. Proc. Natl. Acad. Sci. 104:18461. 18. Tirion, M. M., 1996. Large amplitude elastic motions from single-parameter, atomic anal- ysis. Phys. Rev. Lett. 77:1905 -- 1908. 19. Atilgan, A. R., S. R. Durell, R. L. Jernigan, M. C. Demirel, O. Keskin, and I. Bahar, 2001. Anisotropy of Fluctuation Dynamics of Proteins with an Elastic Network Model. Biophys. J. 80:505 -- 515. 20. Tama, F., and Y. H. Sanejouand, 2001. Conformational change of proteins arising from normal mode calculations. Protein Eng 14:1 -- 6. 21. Tama, F., M. Valle, J. Frank, and r. Brooks, C. L., 2003. Dynamic reorganization of the functionally active ribosome explored by normal mode analysis and cryo-electron mi- croscopy. Proc. Natl. Acad. Sci. USA 100:9319 -- 23. 22. Tama, F., and C. L. Brooks, 2005. Diversity and identity of mechanical properties of icosahedral viral capsids studied with elastic network normal mode analysis. J. Mol. Biol. 345:299 -- 314. 23. Lu, M., B. Poon, and J. Ma, 2006. A New Method for Coarse-Grained Elastic Normal- Mode Analysis. J. Chem. Theory Comput. 2:464 -- 471. Electro-elasticmodel 16 24. Moritsugu, K., and J. C. Smith, 2007. Coarse-Grained Biomolecular Simulation with REACH: Realistic Extension Algorithm via Covariance Hessian. Biophys. J. 93:3460 -- 3469. 25. Lu, M., and J. Ma, 2008. A minimalist network model for coarse-grained normal mode analysis and its application to biomolecular x-ray crystallography. Proc. Nat. Acad. Sci. 105:15358 -- 63. 26. Riccardi, D., Q. Cui, and G. N. Phillips, 2009. Application of elastic network models to proteins in the crystalline state. Biophys. J. 96:464 -- 475. 27. Bahar, I., T. R. Lezon, L.-W. Yang, and E. Eyal, 2010. Global Dynamics of Proteins: Bridging Between Structure and Function. Ann. Rev. Biophys. 39:23 -- 42. 28. Halle, B., 2002. Flexibility and packing in proteins. Proc. Natl. Acad. Sci. 99:1274 -- 1279. 29. Lu, M., and J. Ma, 2005. The Role of Shape in Determining Molecular Motions. Biophys. J. 89:2395 -- 2401. 30. Tama, F., and C. L. Brooks, 2006. Symmetry, Form, and Shape: Guiding Principles for Robustness in Macromolecular Machines. Annu. Rev. Biophys. Biomol. Struct. 35:115 -- 133. 31. Petrone, P., and V. S. Pande, 2006. Can conformational change be described by only a few normal modes? Biophys. J. 90:1583 -- 1593. 32. Lyman, E., J. Pfaendtner, and G. A. Voth, 2008. Systematic multiscale parametrization of heterogeneous elastic network models of proteins. Biophys. J. 95:4183. 33. Hinsen, K., and G. R. Kneller, 1999. A simplified force field for describing vibrational protein dynamics over the whole frequency range. J. Chem. Phys. 111:10766 -- 10769. 34. Miller, B. T., W. Zheng, R. M. Venable, R. W. Pastor, and B. R. Brooks, 2008. Langevin Network Model of Myosin. J. Phys. Chem. B 112:6274 -- 6281. 35. Matyushov, D. V., 2011. Nanosecond Stokes Shift Dynamics, Dynamical Transition, and Gigantic Reorganization Energy of Hydrated Heme Proteins. J. Phys. Chem. B 115:10715 -- 10724. 36. LeBard, D. N., and D. V. Matyushov, 2009. Energetics of bacterial photosynthesis. J. Phys. Chem. B 113:12424 -- 12437. 37. Tozzini, V., 2010. Minimalist models for proteins: a comparative analysis. Quarterly Reviews of Biophysics 43:333 -- 371. 38. Hinsen, K., 2008. Structural flexibility in proteins: impact of the crystal environment. Bioinformatics 24:521 -- 8. Electro-elasticmodel 17 39. Gerek, Z. N., and S. B. Ozkan, 2011. Change in allosteric network affects binding affinities of PDZ Domains: Analysis through Perturbation Response Scanning. PLoS Comput. Biol. 7:e1002154. 40. Ming, D., and M. E. Wall, 2005. Allostery in a Coarse-Grained Model of Protein Dynamics. Phys. Rev. Lett. 95:198103. 41. Gerek, Z. N., O. Keskin, and S. B. Ozkan, 2009. Identification of specificity and promis- cuity of PDZ domain interactions through their dynamic behavior. Proteins 77:796 -- 811. 42. Smith, J. C., 1991. Protein dynamics: comparison of simulations with inelastic neutron scattering experiments. Quat. Rev. Biophys. 24:227. 43. Cametti, C., S. Marchetti, C. M. C. Gambi, and G. Onori, 2011. Dielectric Relaxation Spectroscopy of Lysozyme Aqueous Solutions: Analysis of the delta-Dispersion and the Contribution of the Hydration Water. J. Phys. Chem. B 115:7144 -- 7153. 44. Lamm, G., and A. Szabo, 1986. Langevine modes of macromolecules. J. Chem. Phys. 85:7334. 45. Ansari, A., 1999. Langevin modes analysis of myoglobin. J. Chem. Phys. 110:1774 -- 1780. 46. Erkip, A., and B. Erman, 2004. Dynamics of large-scale fluctuations in native proteins. Analysis based on harmonic inter-residue potentials and random external noise. Polymer 45:641 -- 648. 47. Hansen, J. P., and I. R. McDonald, 2003. Theory of Simple Liquids. Academic Press, Amsterdam. 48. Essiz, S. G., and R. D. Coalson, 2009. Dynamic Linear Response Theory for Conforma- tional Relaxation of Proteins. J. Phys. Chem. B 113:10859 -- 10869. 49. Soheilifard, R., D. E. Makarov, and G. J. Rodin, 2011. Rigorous coarse-graining for the dynamics of linear systems with applications to relaxation dynamics in proteins. J. Chem. Phys. 135. 50. Skepo, M., P. Linse, and T. Arnebrant, 2006. Coarse-Grained Modeling of Proline Rich Protein 1 (PRP-1) in Bulk Solution and Adsorbed to a Negatively Charged Surface. J. Phys. Chem. B 110:12141 -- 12148. 51. Pizzitutti, F., M. Marchi, and D. Borgis, 2007. Coarse-Graining the Accessible Surface and the Electrostatics of Proteins for Protein-Protein Interactions. J. Chem. Theory and Comp. 3:1867 -- 1876. 52. Leherte, L., and D. P. Vercauteren, 2009. Coarse Point Charge Models For Proteins From Smoothed Molecular Electrostatic Potentials. J. Chem. Theory Comp. 5:3279 -- 3298. 53. Dong, F., B. Olsen, and N. A. Baker, 2008. Methods in Cell Biology, Elsevier, volume 84, chapter Computational methods for biomolecular electrostatics, 843. Electro-elasticmodel 18 54. Berardi, R., L. Muccioli, S. Orlandi, M. Ricci, and C. Zannoni, 2004. Mimicking electro- static interactions with a set of effective charges: a genetic algorithm. Chemical Physics Letters 389:373 -- 378. 55. Chaikin, P. M., and T. C. Lubensky, 1995. Principles of condensed matter physics. Cam- bridge University Press, Cambridge. 56. LeBard, D. N., and D. V. Matyushov, 2008. Redox Entropy of Plastocyanin: Developing a Microscopic View of Mesoscopic Solvation. J. Chem. Phys. 128:155106. 57. MacKerell, A. D., D. Bashford, M. Bellott, R. L. D. Jr., J. D. Evanseck, M. J. Field, S. Fischer, J. Gao, H. Guo, D. S. Ha, J. McCarthy, L. Kuchnir, K. Kuczera, F. T. K. Lau, C. Mattos, S. Michnick, T. Ngo, D. T. Nguyen, B. Prodhom, W. E. R. III, B. Roux, M. Schlenkrich, J. C. Smith, R. Stote, J. Straub, M. Watanabe, J. Wirkiewicz-Kuczera, D. Yin, and M. Karplus, 1998. All-atom empirical potential for molecular modeling and dynamics studies of proteins. J. Phys. Chem. B 102:3586 -- 3616. 58. Phillips, J. C., R. Braun, W. Wang, J. Gumbart, E. Tajkhorshid, E. Villa, C. Chipot, R. D. Skeel, L. Kale, and K. Schulten, 2005. Scalable molecular dynamics with NAMD. J. Comp. Chem. 26:1781 -- 1802. 59. Jorgensen, W. L., J. Chandrasekhar, J. D. Madura, R. W. Impey, and M. L. Klein, 1983. Comparison of simple potential functions for simulating liquid water. J. Chem. Phys. 79:926 -- 935. 60. Humphrey, W., A. Dalke, and K. Schulten, 1996. VMD - Visual Molecular Dynamics. Journal of Molecular Graphics 14:33 -- 38. 61. LeBard, D. N., and D. V. Matyushov, 2010. Ferroelectric hydration shells around proteins: Electrostatics of the protein-water interface. J. Phys. Chem. B 114:9246 -- 9258. 62. Autenrieth, F., E. Tajkhorshid, J. Baudry, and Z. Luthney-Schulten, 2004. Classical force field parameters for the heme prosthetic group of cytochrome c. J. Comp. Chem. 25:1613. 63. Meuwly, M., O. M. Becker, R. Stote, and M. Karplus, 2002. NO rebinding to myoglobin: a reactive molecular dynamics study. Biophys. Chem. 98:183. 64. Zhang, Y., and J. E. Straub, 2009. Diversity of solvent dependent energy transfer pathways in heme proteins. J. Phys. Chem. B 113:825. 65. Gabel, F., D. Bicout, U. Lehnert, M. Tehei, M. Weik, and G. Zaccai, 2002. Protein dynamics studied by neutron scattering. Quat. Rev. Biophys. 35:327 -- 367. 66. Zaccai, G., 2000. How soft is a protein? A protein dynamics force constant measured by neutron scattering. Science 288:1604 -- 1607. 67. Weiss, S., 2000. Measuring conformational dynamics of biomolecules by single molecule fluorescence spectroscopy. Nat. Struct. Biol. 7:724 -- 729.
1608.04824
1
1608
2016-08-17T01:14:43
Noise and Function
[ "physics.bio-ph", "math.PR", "nlin.AO", "q-bio.QM" ]
Noise is widely understood to be something that interferes with a signal or process. Thus, it is generally thought to be destructive, obscuring signals and interfering with function. However, early in the 20th century, mechanical engineers found that mechanisms inducing additional vibration in mechanical systems could prevent sticking and hysteresis. This so-called "dither" noise was later introduced in an entirely different context at the advent of digital information transmission and recording in the early 1960s. Ironically, the addition of noise allows one to preserve information that would otherwise be lost when the signal or image is digitized. As we shall see, the benefits of added noise in these contexts are closely related to the phenomenon which has come to be known as stochastic resonance, the original version of which appealed to noise to explain how small periodic fluctuations in the eccentricity of the earth's orbit might be amplified in such a way as to bring about the observed periodic transitions in climate from ice age to temperate age and back. These noise-induced transitions have since been invoked to explain a wide array of biological phenomena, including the foraging and tracking behavior of ants. Many biological phenomena, from foraging to gene expression, are noisy, involving an element of randomness. In this paper, we illustrate the general principles behind dithering and stochastic resonance using examples from image processing, and then show how the constructive use of noise can carry over to systems found in nature.
physics.bio-ph
physics
Noise and Function Steven Weinstein∗ Theodore P. Pavlic† Abstract Noise is widely understood to be something that interferes with a signal or process. Thus, it is generally thought to be destructive, obscuring signals and interfering with function. However, early in the 20th century, mechanical engineers found that mecha- nisms inducing additional vibration in mechanical systems could prevent sticking and hysteresis. This so-called "dither" noise was later introduced in an entirely different context at the advent of digital information transmission and recording in the early 1960s. Ironically, the addition of noise allows one to preserve information that would otherwise be lost when the signal or image is digitized. As we shall see, the benefits of added noise in these contexts are closely related to the phenomenon which has come to be known as stochastic resonance, the original version of which appealed to noise to explain how small periodic fluctuations in the eccentricity of the earth's orbit might be amplified in such a way as to bring about the observed periodic transitions in climate from ice age to temperate age and back. These noise-induced transitions have since been invoked to explain a wide array of biological phenomena, including the foraging and tracking behavior of ants. Many biological phenomena, from foraging to gene ex- pression, are noisy, involving an element of randomness. In this paper, we illustrate the general principles behind dithering and stochastic resonance using examples from image processing, and then show how the constructive use of noise can carry over to systems found in nature. 1 Introduction We are surrounded by noise. The controlled explosions of internal combustion engines combine with the roar of rubber on asphalt to create the drone of road and highway traffic. Weed-whackers, lawnmowers, and leaf blowers can turn sunny suburban summer days into a buzzing confusion. Indeed, the entire universe is filled with the faint din that is the cosmic microwave background (CMB) radiation, leftover electromagnetic radiation from the The CMB was discovered when Penzias and Wilson (1965) went looking for the source of annoying hiss plaguing their new radio telescope, hiss that threatened to obscure signals from distant stars and galaxies. Noise seems to be entirely destructive, thus something to be eliminated if possible. Noise can be beneficial, however, in at least two ways. One is familiar, the other paradox- ical and far less well known. The familiar way is simply as a source of variety. For example, genes undergo random mutation from processes both external to the organism (e.g. cosmic rays) and internal (Dobrindt and Hacker, 2001). This genotypic variation is the source of ∗Department of Philosophy, University of Waterloo, Waterloo, ON, Canada N2L 3G1 †School of Computing, Informatics, and Decision Systems Engineering, Arizona State University, Tempe, AZ, USA 85287 1 heritable variation in phenotype, which is of course essential for the process of natural selec- tion (Wagner, 2014). Phenotype can even vary within isogenic populations due to variation in gene expression (Fraser and Kaern, 2009). This phenotypic noise is thought to provide an evolutionary advantage for some microorganisms, as it increases the chance that some will survive under stressful conditions. The CMB noise, though destructive from the standpoint of the users of radio telescopes, plays a constructive role in generating the tiny variations in energy density in the early universe that are the seeds of structure formation. The random fluctuations we call noise give rise to stars and galaxies and galactic clusters. This much is familiar, at least to those working in the relevant areas of biology or astrophysics. Considerably less familiar is the role that noise can play in nonlinear systems, in par- ticular systems with one or more thresholds, points at which small differences in input give rise to disproportionate differences in output. Converting an analog signal into a digital signal involves sampling the signal at regular intervals and writing down a digital approx- imation of the amplitude at each point. For example, if one has a one-bit digital system with two possible values to represent the interval from 0 to 1, then there will be a threshold at the analog value 0.5, below which any value will be digitally recorded as 0, and above which any value will be recorded as 1. More bits simply mean more thresholds, more ways to cut up the interval into discrete chunks. Neurons behave like single threshold devices, firing when and only when the voltage across the cell membrane reaches a certain activation threshold. What noise can do in a threshold system is push the signal over the threshold, but in a way very much unlike an amplifier. Amplifiers multiply the signal, whereas noise is additive. The implications and applications of this nonstandard amplification are both deep and wide. Here we will lay out as simply as possible the principles behind this sort of noise benefit and then illustrate its application. 2 Shades of gray: Noise in image processing Photographs are never veridical. The information coming through the lens is inevitably greater than the information stored on the recording medium. A digital camera sensor has a finite spatial resolution; the camera's sensor consists of a matrix of smaller individual sensors corresponding to a single "pixel" of the image. Any features of the image smaller than an individual pixel will be lost. Associated with each pixel is a color. The color spectrum in the real world is continuous, but the digital encoding of color is discrete, so that in general, the color stored will only be an approximation of the actual color. In other words, color information must be rounded off in order to be stored as a number on a digital computer. The number of binary places available for each number is referred to as the bit depth. Suppose we have a digital image consisting of 7 megapixels. Let's consider a "black and white" camera for simplicity, so that the colors are shades of gray. Each pixel has an 8 bit number attached to it indicating the shade of gray, with 0 (00000000) corresponding (by convention) to black, and 255 (11111111) corresponding to white. There are a total of 28 = 256 shades of gray. Figure 1(a) shows the palette, alongside an image of a woman known as Lena (Hutchison, 2001), rendered using this palette (Figure 1(b)). Now, suppose we want to print these images. Indeed, you may well be reading a printed version of this page, printed on a laser printer capable of printing 300 dots per inch (dpi). That resolution gives us around 7 million evenly spaced points on a typical sheet of paper, so if we were to use up an entire page to print the 7 megapixel image, we would have a one- 2 3 (a) 8 bit grayscale palette (256 shades) (b) 8 bit Lena (256 shades) (c) 1 bit grayscale (2 shades) (d) 1 bit Lena (2 shades) (e) 3 bit grayscale (8 shades) (f) 3 bit Lena (8 shades) Figure 1: Shades of gray (a) Halftone array (b) Halftone grayscale palette Figure 2: Grayscale using halftone to-one correspondence between pixels and dots. If the printer could print 256 shades of gray at any given point, then we'd have perfect reproduction of the stored image. But the printer is not nearly that flexible. At each dot location, most printers can print either a black dot or nothing. Because the number of pixels and the number of dots are approximately the same (in our example), we are effectively reducing an 8-bit (per pixel) image to a 1-bit (per dot) image; 256 shades of gray at each point get mapped to either black or white. The obvious way to map the shades of gray is to impose a threshold as before, whereby we print a black dot at a point if the corresponding pixel is more than 50% gray (numbers between 0 and 127), and we otherwise leave it blank (white) (numbers between 128 and 255). For an image that has an equal distribution of lighter and darker grays, this might appear to be as good as one can do. But a quick glance at Figures 1(c) and 1(d) shows the limita- tions of this simple thresholding method; an enormous amount of detail is lost. Increasing the number of thresholds to create 8 shades of gray, as in Figures 1(e) and 1(f), yields a noticeable improvement, but a printer that can only print in black and white is limited to the performance of the 2 shade case. However, there are clever methods to improve the fidelity of black-and-white image reproductions. Traditional printed newspapers used varying dot size to represent darker and lighter portions of an image. Applying this halftone concept to a device like a laser printer or an LCD display with fixed dot or pixel size involves representing gray by varying the density of the distribution of black dots in an array. Using 3 × 3 arrays of dots, we can represent ten different shades of gray shown in Figure 2(a), which allows for a grayscale palette like that shown in Figure 2(b). A 300 dpi (dots per inch) printer can print 100× 100 patches of gray per inch, where each patch has a 3 × 3 pixel area. Thus, with a three-fold reduction in the effective spatial resolution of the image, the number of shades which can be represented is increased from two shades to ten. The same technique could be applied to 4 × 4 arrays of dots to achieve seventeen shades of gray at the cost of further decreasing the spatial resolution. 4 Although deterministic methods like halftoning can be effective ways of trading spatial resolution for color resolution, they introduce noticeable artifacts. Note the abruptness of the shifts in the halftone grayscale palette in Figure 2(b). To avoid this blockiness while still being able to trade spatial resolution for color resolution, a very different approach can be used based on the appropriate addition of random variation to pixel values. For each pixel, we take the original grayscale value and add a random number between 0 and 255. An image made up of these random values would look like visual "noise" – it is a distribution of dots in arbitrary shades of gray. The result of adding this noise to the image is an array of pixels with values between 0 and 510. We can turn the resulting array back into an image by dividing the values by two, thereby restoring the original range of 0 (black) to 255 (white). Figures 3(a) and 3(b) show the result: noisy versions of the original grayscale palette and the original image. This of course is not an improvement over the non-noisy 8-bit grayscale image with 256 shades of gray. The noise does what we generally expect noise to do: it degrades the image. But recall that we added the noise not to improve the grayscale image but to get a better result when we subsequently impose a threshold at 127 and convert to a black and white (2 shade) image. Imposing the threshold, we map any pixel with value 128 to white (255), and any pixel 127 or below to black (0). The resulting pseudo-grayscale palette that looks like Figure 3(c), while the resulting pseudo-grayscale Lena looks like Fig- ure 3(d). The results are instructive. The grayscale palette looks better than the version in which noise was not imposed before thresholding (Figure 1(c)), giving the impression of a variety of shades of gray. Lena, however, does not look very good by comparison with Figure 1(d). The reason, as we noted above, is that given enough pixels, the ratio of black dots to white will closely approximate the degree of grayness of the original shade from the 256-color palette. For a large number of pixels, observed at a sufficiently great distance, we get an excellent representation of gray. The problem with the Lena image is not that she has more shades of gray, but that the shades tend to change over the scale of a few pixels. Images of this sort are better treated by more sophisticated techniques such as the Floyd– Steinberg error diffusion method (Floyd and Steinberg, 1976). However, if we avail ourselves of 8 shades of gray (3 bits) rather than just black and white (1 bit), the use of random noise is much more effective. Figure 4(a) and Figure 4(b) show the original 256 shade images augmented with a low level of noise, spanning 1/8 of the total range (32 shades of gray). That is, the noise randomizes the 3 least significant bits (LSB) of the 8 bits in use. If we now reduce to to 8 shades (encodable by 3 bits) having added this noise, we get Figure 4(c) and Figure 4(d), which are a decided improvement on Figure 1(e) and Figure 1(f), which are what we get if we go from 256 shades to 8 shades without first adding noise. The process of adding noise to an image or a signal in order to preserve information once the signal is subjected to quantization1 (digitization) is called dithering.2 In the example above, we took an already discrete signal (each pixel having one of 256 shades of gray) and made it more discrete, mapping the 256 shades into 2 shades (black and white). However, the initial process of moving from an image with a continuum of shades to one with 256 shades is also an example of quantization. Were we dealing with digital audio, we would 1"Quantization" here does not refer to the physicist's process of finding a quantum-mechanical version of a classical theory, but rather the process of discretizing the properties of an image or signal. 2Dithering also includes related methods that use, not random noise, but some other signal which is uncorrelated with the signal of interest. The ring laser gyroscope, for example, uses periodic (sinusoidal) dither to prevent its counter-rotating laser beams from locking under conditions of slow rotation. 5 6 (a) 8 bit grayscale (256 shades) w/ 8 bit noise (b) 8 bit Lena (256 shades) w/ 8 bit noise (c) 1 bit grayscale (2 shades) (d) 1 bit Lena (2 shades) Figure 3: Reduction from 256 shades (8 bits) to 2 shades (1 bit) using 8 bit random dither noise. 7 (a) 8 bit grayscale (256 shades) w/ 3 bit noise (b) 8 bit Lena (256 shades) w/ 3 bit noise (c) 3 bit grayscale (8 shades) (d) 3 bit Lena (8 shades) Figure 4: Reduction from 256 shades (8 bits) to 8 shades (3 bits) using 3 bit random dither noise (randomizing the 3 least significant bits (LSB). be working with a one-dimensional stream of samples of the waveform, each of which has a continuously valued amplitude (the volume) that must be mapped into a finite set of numbers for storage in a computer, say 24 bits (16,777,216 possible values). This can then be further reduced to 16 bits (64,436 possible values) for CD encoding. Dither is routinely used in this process. Most digital representations involve more than one threshold. The seminal work of Roberts (1962) considered the problem of transmitting digital television. At the time, a 6-bit-per-pixel resolution was considered adequate. Roberts proposed a scheme whereby a 3- bit-per-pixel signal could effectively encode the necessary detail if pseudo-random noise were added to the 6-bit representation prior to rounding to 3 bits.3. Shortly thereafter, others realized that this technique was akin to a technique called "dither" that had been conceived several decades prior as a technique to overcome the tendency of certain mechanical systems to stick for various reasons, rendering them insensitive to small changes in operational parameters (Schuchman, 1964). Engineers designed electromotor circuits to apply dither in the form of small zero-mean oscillations that would allow devices to respond more easily to small steering signals from the operator (Farmer, 1944; Korn and Korn, 1952). In all of these cases, proper application of dither tends to linearize a nonlinear system; the dither blurs thresholds, eliminating some of the jaggedness that goes along with systems that have one or more thresholds. But the addition of noise does something else along the way. In a physical system, adding noise adds energy to the system. If the system has one or more thresholds, this has the effect of taking subthreshold signals and boosting them, albeit stochastically. This is the phenomenon known as stochastic resonance. Let's take a look at a simple example, again using printed images, before moving on to the role stochastic resonance can play in natural – including biological – phenomena. We're accustomed to the fact that there are sounds we can't hear because they're too soft, and sights we can't see because they're too faint. These are thresholds of hearing and vision, respectively. By analogy, consider an image we can't make out because it's too light: a very light shade of gray indistinguishable from its white background. Figure 5(a) shows the full grayscale palette with a line separating the grays that are dark enough to distinguish from those that, for some focal individual, are not. We can represent the indistinguishability of any grays below the threshold by rendering them as white, as in Figure 5(b). Thus, when a faint image of the words 'Phantom Engineer' (Figure 5(c)) is rendered in this very light gray, it will look like Figure 5(d) to someone for whom this threshold represents the limits of their perceptual acuity. The image will be invisible. Now, suppose we add noise by randomly darkening each pixel, including the background. We will use low-level, 3 bit noise, as shown to the right of the vertical bar in the grayscale palette depicted in Figure 6(a). This corresponds to randomly selected shades of the very light grays lying below the threshold of perceivability. Thus, like the image itself, the noise will be invisible – see Figure 6(b) – to someone who cannot make out very light shades of gray. Adding this noise to our original image as in Figure 6(c) has the remarkable effect of bringing a noisy but very legible version of the image above the threshold of perceivability, as is evident in Figure 6(d), in which the lightest, sub-threshold grays are removed. The ability of noise to boost a signal above threshold is the essence of stochastic resonance. One of the salient characteristics of stochastic resonance, indeed the feature that makes it somewhat akin to a true resonance phenomenon, is the dependence on the amplitude and 3The scheme of Roberts (1962) was an early example of what is called subtractive dither, where the noise is subtracted from the image after transmission 8 9 (a) 8 bit grayscale before thresholding (threshold marked) (b) 8 bit grayscale after thresholding (c) 8 bit signal before thresholding (d) 8 bit signal after thresholding Figure 5: Rendering a signal imperceivable by thresholding. The signal is too light to survive the imposition of the threshold. 10 (a) Low-level (3 bit) noise before thresholding (b) Low-level (3 bit) noise after thresholding (c) 8 bit signal plus 3 bit noise before thresholding (d) 8 bit signal plus 3 bit noise after thresholding Figure 6: Stochastic resonance: signal boosting with noise. Random dither noise spanning the 3 least significant bits (LSB) is added. With the addition of noise, the signal becomes dark enough to remain visible even after the imposition of the threshold. specific properties of the noise. This is true of dither noise in general. Too much noise (here, too large a spectrum of grays) threatens to obliterate the signal, while too little will fail to push the signal above threshold at all, and have no effect. The relevance of this for understanding the role of stochastic resonance in nature is significant, for there is noise of all kinds and all amplitudes everywhere. Oftentimes it does what we think noise does: it interferes with the signal, the image, or the operation of a dynamical system. But when the noise level is proportional to the level of one or more significant thresholds in a system of interest, we can and should look for stochastic resonance, as it may be key to understanding the function of the system (Gammaitoni et al., 1998). 3 Amplification by noise in natural dynamical systems We will now take a look at how the stochastic resonance effect can be used in modeling a dynamical system existing in nature. The example we'll study is the one in which the term 'stochastic resonance' was originally introduced. Though there is no resonance in the ordinary physicists' sense (though see Gammaitoni et al., 1995), we are once again presented with a situation in which the addition of noise permits the system of interest – in this case the climate – to straddle a threshold. The earth has existed in two relatively stable climates around 10 degrees Kelvin apart for millions of years. Periods in which the climate is cooler are called ice ages. In the late 1970s (Bhattacharya and Ghil, 1978), it was conjectured that the two stable climates corre- spond to the two-minima of a double-well pseudo-potential like the one shown in Figure 7. The horizontal axis represents the earth's temperature, and the curve acts like a potential energy term in ordinary mechanics, with a single unstable equilibrium forming an energy barrier between two stable equilibria. In this model, the earth's climate inevitably converges to one of the two stable equilibria. The equilibrium on the left represents the ice age, and the one on the right is a temperate period like the present. Thus, we have a primitive model of a system with two stable states. But the stability of these states means that there is no way to transition between them, thus no way to explain how the climate shifts from one to the other. However, it was observed that the eccentricity of the earth's orbit varies over a period roughly the same as the time between ice ages – around 100,000 years (Hays et al., 1976). Because the eccentricity of the orbit is correlated with small variations in the amount of solar heating ("insolation"), it was conjectured that these small variations might be sufficient external drivers of the earth's dynamical system to cause the observed periodic climate changes. This suggests that one augment the model by introducing a time-varying oscillation in the pseudo-potential in which the double-well shape is a transient feature separating two epochs in which the potential morphs into a single well, a single quasi-stable equilibrium. The problem with this idea was that the estimated insolation differences were too small to be responsible for such a change. At best, the resulting time-varying pseudopotential takes forms like those in Figure 8, where the barrier between the two stable equilibria is always maintained, and where the change in temperature due to the displacement of each local minimum is of the order of only 1 degree Kelvin. So a simple dynamical systems approach does not provide an explanation for the congruence of the period of the insolation signal and the period of earth's climate switching. Independently, Nicolis (1982) and Benzi et al. (1982) arrived at similar explanations for how the climate might actually shift. They proposed that including the fine-scale, shorter- 11 Figure 7: Double-well pseudo-potential representing the earth's climate as a dynamical system time variations in heating and cooling due to various other factors might result in the climate hopping from one well to the other. In other words, factoring in the existence of a certain level of noise in the climate system might account for the ability, as it were, of the climate to surmount the otherwise insurmountable threshold, the hump between the two minima. After all, the geological record shows not only periodicity in the earth's climate but also significant small-scale variations which indicated that the earth's dynamics must include some internal noise. So, following the approach of Nicolis, the deterministic double- well pseudo-potential is augmented with a noise source, converting an ordinary differential equation into a stochastic differential equation. In other words, the climate is now modeled as a diffusion process – a random walk that is pulled downhill but can, on occasion, take several steps uphill. For such a diffusion process in a double-well, the mean time to transition from one stable equilibrium to another is well characterized by a formula parameterized by the height of the barrier between the equilibria and the strength of the internal noise. What Nicolis realized was that the changes in the barrier height due to noise could lead to large changes in the mean residence time. If the noise is of the right amplitude, then the climate is likely to hop from one not-quite-stable minimum to the other when the barrier is low. Consequently, Benzi et al. (1982) named the phenomenon stochastic resonance based on its similarity to the frequency-selective properties of conventional deterministic resonance. Whereas "resonance" in the traditional sense is between the frequency of an input and the characteristic response of a system, the resonance here is between the frequency of the long- term oscillation in insolation (the input) and the amplitude of the noise (a characteristic feature of the system). If the noise is too small, nothing special will happen; the system will 12 13 (a) t = 0 (b) t = 25, 000 years Figure 8: Time-varying double-well pseudo-potential with period of 100,000 years (c) t = 50, 000 years never transit from one climate to the other. If the noise is too great, the system will never settle in one climate or another, as the noise will dominate the oscillation. If the noise is within the correct range, however, the climate will oscillate at approximately the 100,000 year period of the subthreshold background oscillation in the ellipticity of the earth's orbit. 4 Noise in living systems: decision making in ant colonies Over the past 20 years, an intriguing body of evidence has pointed to a role for stochastic resonance in a variety of biological processes (McDonnell and Abbott, 2009). Extensive work has been done demonstrating the role of noise in general and stochastic resonance in particular in the neural systems that carry out sensory information processing (Moss et al., 2004), but it is important at the macroscopic level as well. We will conclude our discussion of stochastic resonance by discussing its role in the social dynamics of group decision making in certain species of ants. There are a wide variety of mass-recruiting ants that form charismatic foraging trails that concentrate all foraging effort onto a single food source for a short period of time (Holldobler and Wilson, 1990). As many species of mass-recruiting ant have a heterogeneous foraging force, it is thought that it may be beneficial to concentrate the foraging force all in one area in order to guarantee there is an adequate representation of each worker type. So it is expected that these ants must make use of some decentralized mechanism that can drive its foraging force to a quick consensus on the best of several available foraging options. A typical feature of mass-recruiting ants is the use of pheromone trails (Holldobler and Wilson, 1990). Although details vary across different mass-recruiting ant taxa, the observed pattern is usually a variation of what follows. A focal ant leaves her nest and searches for food. When she finds food, she can choose to return to her nest and deposit some quantity of pheromone along her path back to the nest. The amount of pheromone she deposits is related to the quality of the discovered food, with higher-quality foods leading to more deposited pheromone. Although that deposited pheromone will eventually evaporate, for a short time after deposition, the pheromone near the nest will attract the attention of other foragers that would otherwise search randomly for food. They will then have an increased likelihood of finding the same food source as the focal ant and then also lay a pheromone trail on their return visit. So initially, a set of food items will be discovered randomly. Due to the positive feedback inherent in the recruitment system, the highest quality of those food sources will eventually attract all of the foragers. Until recently, it has been believed that such trail-laying mass-recruitment mechanisms had a flaw similar to the one described in Section 3 for the early deterministic models of cli- mate change – the ants were thought to be rigidly bistable and unable to cope with changes in food availability after a critical point in the recruitment process. In other words, the pos- itive feedback in the recruitment would eventually become so strong that the system would become entirely insensitive to changes in relative food-source quality, just as early mathe- matical models of climate change were not properly sensitive to the variations in insolation. This intuition was verified in early experiments with Lasius niger (Beckers et al., 1990). Moreover, early dynamical mathematical models of trail-laying were also shown to be insen- sitive to changes in relative food quality (Camazine et al., 2001; Nicolis and Deneubourg, 1999). However, several recent experiments show that many other trail-laying ants are able to dynamically re-allocate their foraging forces to track changes in the environment (Dussu- tour et al., 2009; Latty and Beekman, 2013; Reid et al., 2011). For example, Dussutour et al. 14 Figure 9: Graphical summary of dynamic foraging experiment used by Dussutour et al. (2009) to study flexibility of trail-laying in Pheidole megacephala big-headed ant colonies. In the first 60 minutes of the experiment, colonies are given a choice between two feeders, A and B, that only differ in distance to the nest. During the second 60 minutes of the experiment, the nearest feeder (A) is removed. Finally, during the final 60 minutes of the experiment, the nearest feeder (A) is replaced. (2009) presented colonies of Pheidole megacephala with a laboratory dynamic environment summarized in Figure 9. During the 180-minute experiment, colonies were placed at the mouth of a Y-bridge with two legs of different lengths, and the experiment proceeded in three 60-minute phases: • During the first 60 minutes, equal-quality feeders were placed at the ends of both legs. Because one leg was shorter, it eventually dominated the collective attention of the colony and a single trail was formed to the feeder on that leg. • During the second 60 minutes of the experiment, the feeder on the short leg was removed. With the disappearance of food, the pheromone trail was not reinforced, and the colony was eventually able to return to random search and subsequently converge on the feeder at the end of the long leg. • During the final 60 minutes of the experiment, a feeder was returned to the short leg of the Y-maze. The traditional model of trail-laying recruitment would predict that the new feeder would be ignored because all foragers would be latched into following the existing pheromone trail. However, contrary to those predictions, the short leg was re-discovered and ants returned to exploiting the closer feeder. To explain the results of the experiment, Dussutour et al. propose a slight extension to the traditional mathematical model of trail-laying recruitment inspired by stochastic reso- nance. They observed that ants in their experiments would often make "errors" in their trail-following behavior that would lead a minority of the ants down the opposite leg of the Y-maze. In an attempt to capture this phenomenon, they augmented the traditional mathematical model of trail-laying behavior with an "error" level that would cause an in- dividual to rarely, but measurably often, choose the leg of the Y-maze with the smaller quantity of pheromone deposited on it. At small error levels, the theoretical system had 15 bbNestB60minutesANestB60minutesNestB60minutesARemoveFeederAReplaceFeederA nearly identical decision-making latency and accuracy characteristics to the deterministic system when presented with a static choice set. However, at specific non-zero error levels, the system could produce switching dynamics that matched those of experimental data from ants like P. megacephala that have the ability to follow changes in relative feeder quality. Furthermore, the mathematical model predicted that different error levels would correspond to different random natural switching times between alternatives, and amplification of vari- ations in food quality would be possible if the periodicity of those variations matched the natural error-driven switching time. The time-scale matching argument for the switching behavior observed in some trail- laying, mass-recruiting ants is identical to the one used in the early models of stochastic resonance in climate systems that are matched to the periodicity of solar insolation. How- ever, in the case of the ants, differences in error level across different ant taxa could be explained by natural selection. In particular, the individual error level could be tuned by natural selection so that the stochastic switching time of the colony would match the natural periodicity of changes in food quality in the natural environment. Colonies with individ- uals that make errors at the appropriate rate would have an advantage over colonies that make more or fewer errors. Making too many individual-level errors would mean switch- ing too frequently from good choices to bad choices, and making too few individual-level errors would mean focusing for too long on one choice even though a better choice was now available. Consequently, the "errors" at the individual level would be better described as random variability (noise) that was itself a trait under selection, and the prediction would be that ants evolved in more ephemeral environments would also have larger amounts of noise in individual-level response to pheromone trails. In fact, as Dussutour et al. (2009) discuss, the flexible P. megacephala ants in their study that are well modeled by non-zero noise do come from an environment where food quality changes more frequently than the L. niger ants that had previously been used to support the deterministic modeling of trail- following behavior with no noise. If location of the best-quality food source is viewed as a signal that tends to change over some characteristic time scale, then the current location of the main foraging trail can be viewed as a version of that signal amplified using stochastic resonance. This argument is identical to those made in the stochastic-resonance literature where some input-to-output measure, such as mutual information, is maximized by varying the amount of noise added to the input signal (Neiman et al., 1996). Similarly, in the earlier image processing example, a certain amount of noise is sufficient to push the text 'Phantom Engineer' above the threshold of visibility (Figure 6). Too little noise will not do the trick. Adding noise will make the text more visible up to a point, after which the readability goes down as the entire image becomes dominated by noise. In the case of the ants, the signal being amplified is the relative quality of the feeders, and the output is the selection of a path by the colony. The idea that apparent "errors" could actually be an adaptive phenotypic trait under selection is not unprecedented and goes beyond the examples of possible stochastic reso- nance in natural phenomena. For example, the idea that noise can be of benefit in decision making is relatively old. For example,in what they called "ethological cybernetics," Haldane and Spurway (1954) performed an information-theoretic analysis of the statistical distribu- tions of honeybees responding to communicated information from so-called "waggle dances." Honeybees have the ability to communicate information from one forager to another through a dance language that communicates the relative polar coordinates (i.e., distance and di- rection) of a discovered food source. However, after a bee communicates these coordinates to another, the bee receiving the information will often make "mistakes" and explore a lo- 16 cation slightly different from the one discovered by the original bee. The average location explored by an ensemble of receiver bees will closely match the originally discovered food source, and so these variations are viewed as "noise" due to imperfect communication of the coordinates in the dance-communication channel. Haldane and Spurway determine that the bee-to-bee channel communicates roughly 4 bits of information about the direction of the target. That is, a bee can only communicate 1 of 16 different cardinal directions; any finer resolution appears to be impossible. This may seem to reflect a fundamental limitation, such as a physiological or neurological constraint, but it could also be an adaptive response to dispersed food sources in an environment. If a honeybee is dancing to communicate the location of a nectar source, such as a flower, the resulting noisy scatter of her colony mates will likely find other flowers in a similar location. Thus, the amount of error in the communication may be tied to some ecological measure of forage patchiness, and honeybees selected for environments with a different patchiness may communicate with different levels of error. As the honeybee example does not involve dynamically changing signals in the envir- onment, it is not an example of stochastic resonance in the strict sense, but it does reflect how the amount of noise expressed in a behavior is itself a phenotype that nature can adapt to match natural variation. However, a very similar information-theoretic analysis of fire ants does suggest additional ties to stochastic resonance and behaviors shaped by nature. In particular, Wilson (1962) described a consistent error distribution in the distance and di- rection information communicated by fire-ant pheromone trails leading to prey. In the trail- laying examples above, the actual food sources were static, and the experimenters could change the location of different sources at discrete instants of time. This was appropriate for the particular ant species under study. However, fire ants are a natural example of a species adapted to continuously varying food quality and location. These ants track moving food sources: living prey items that have the ability to flee. They must have the ability to dynamically adapt and follow fleeing prey until the prey is sufficiently subdued. Wilson suggests that the relatively poor ability of individual fire ants to follow trails is actually an adaptation. The result of the ensemble of error-prone trail followers is a cloud of ants in the general vicinity of the original location of the discovered prey. If this cloud is large enough, it can track the motion of the escaping prey. Too much noise will cast too large of a net and lead to too thin coverage over a prey item, and too little noise will not disperse the ants far enough to catch the escaping prey. So the dispersal of the trail followers could be matched to the escape dynamics of the typical kinds of prey. Just as in traditional stochastic resonance examples, a certain critical amount of noise helps a dynamic output (the ultimate location of the end of a fire-ant foraging trail) follow a dynamic input (the trajectory of an escaping prey item). 5 Conclusion An ant colony's ability to react to changes in the food supply and the climate's ability to react to small changes in insolation are examples of the power of noise to qualitatively change the way a system responds to its environment. If we disregard the noise – disregard the small, random variations in the properties of the system – we find the system converges to a fixed point of its dynamical equations and stays there indefinitely. The ant colony remains fixated on a single food source, insensitive to changes in food supply; the climate remains where it is, never shifting. But when the model of the system is modified to include 17 noise, the system is able to surmount a barrier and transition to a qualitatively distinct state. The colony is able to discover and consolidate a new path. The climate reacts to the slight change in insolation over the course of millenia. Noise makes these systems more sensitive to their environment. A promising area to look for further noise effects in biology lies at smaller scales. The fundamental process at the foundation of all life is the expression of genes as proteins. The chemical reactions involved are constrained by both the availability of the reactants and their proximity, making the process as a whole subject to fluctuations which have come to be called gene expression noise (Kaern et al., 2005). The sources of the noise and the role it plays in the process of development and reproduction are matters of intense contemporary investigation (Sanchez et al., 2013; Viney and Reece, 2013). For example, Fernando et al. (2009) proposed the existence of an intracellular genetic perceptron, a single cell gene network capable of associative learning. Remarkably,Bates et al. (2014) show that the performance of the perceptron is actually enhanced by gene expression noise at a specific level. One of the factors impacting gene expression in bacteria is the intercellular commu- nication mechanism broadly known as quorum sensing, in which bacteria both emit and detect signaling molecules, allowing them to infer the concentration of other bacteria and act accordingly (Popat et al., 2015; Waters and Bassler, 2005). This, too, is a noisy process, subject to the whims of diffusion in the intercellular environment. Like the examples of stochastic resonance we have considered, it is a threshold-oriented system, whereby genes are switched on or off depending on whether a critical density of other bacteria are sensed in the neighborhood. Karig et al. (2011) have applied stochastic resonance to the development of a synthetic biological system which utilizes gene expression noise to boost a time-varying molecular signal consisting of varying concentrations of the molecule used by Gram-negative bacteria in quorum sensing. The recent, fascinating work on the role in biological systems of stochastic resonance in particular, and noise in general, is surely the tip of an iceberg. Hoffmann (2012) advances the idea that the molecular machines like kinesin that do physical work within the cell make use of the random noise that is the thermal motion of the water molecules in the cytoplasm. Rolls and Deco (2010) provide an extended look at the role of noise in brain function. The idea that noise can and does do work, enhancing the information processing that is essential to life, is an idea whose time has come. References Bates, Russell, Blyuss, Oleg, and Zaikin, Alexey. 2014. Stochastic resonance in an intracel- lular genetic perceptron. Physical Review E, 89(3), 032716. Beckers, R., Deneubourg, Jean-Louis, Goss, Simon, and Pasteels, Jacques M. 1990. Collec- tive decision making through food recruitment. Insectes Sociaux, 37(3), 258–267. Benzi, Roberto, Parisi, G., Sutera, Alfonso, and Vulpiani, Angelo. 1982. Stochastic reso- nance in climate change. Tellus, 34, 10–16. Bhattacharya, K., and Ghil, M. 1978. An energy-balance model with multiply-periodic and quasi-chaotic free oscillations. Pages 299–310 of: Evolution of planetary atmospheres and climatology of the earth. Toulouse, France: Centre National dEtudes Spatiales. 18 Camazine, Scott, Deneubourg, Jean-Louis, Franks, Nigel R., Sneyd, James, Theraulaz, Guy, and Bonabeau, Eric. 2001. Self-Organization in Biological Systems. Princeton, NJ, USA: Princeton University Press. Dobrindt, Ulrich, and Hacker, Jorg. 2001. Whole genome plasticity in pathogenic bacteria. Current opinion in microbiology, 4(5), 550–557. Dussutour, Audrey, Beekman, Madeleine, Nicolis, Stamatios C., and Meyer, Bernd. 2009. Noise improves collective decision-making by ants in dynamic environments. Proc. R. Soc. B, 276(1677), 4353–4361. Farmer, William C. (ed). 1944. Ordnance Field Guide: Restricted. Military Service Pub- lishing Company. Fernando, Chrisantha T, Liekens, Anthony ML, Bingle, Lewis EH, Beck, Christian, Lenser, Thorsten, Stekel, Dov J, and Rowe, Jonathan E. 2009. Molecular circuits for associative learning in single-celled organisms. Journal of the Royal Society Interface, 6(34), 463–469. Floyd, Robert W., and Steinberg, Louis. 1976. An adaptive algorithm for spatial grey scale. Proc. Soc. Inf. Disp., 17, 75–77. Fraser, Dawn, and Kaern, Mads. 2009. A chance at survival: gene expression noise and phenotypic diversification strategies. Molecular microbiology, 71(6), 1333–1340. Gammaitoni, L, Marchesoni, F, and Santucci, S. 1995. Stochastic resonance as a bona fide resonance. Physical review letters, 74(7), 1052. Gammaitoni, Luca, Hanggi, Peter, Jung, Peter, and Marchesoni, Fabio. 1998. Stochastic resonance. Reviews of modern physics, 70(1), 223. Haldane, John B. S., and Spurway, H. 1954. A statistical analysis of communication in "Apis mellifera" and a comparison with communication in other animals. Insectes Sociaux, 1(3), 247–283. Hays, J. D., Imbrie, John, and Shackleton, N. J. 1976. Variation in the earth's orbit: pacemaker of the ages. Science, 194(4270), 1121–1132. Hoffmann, Peter M. 2012. Life's ratchet: how molecular machines extract order from chaos. Basic Books. Holldobler, Bert, and Wilson, Edward O. 1990. The Ants. Harvard University Press. Hutchison, Jamie. 2001. Culture, communication, and an information age Madonna. IEEE Prof. Commun. Soc. Newsl., 45(3), 1, 5–7. Kaern, Mads, Elston, Timothy C., Blake, William J., and Collins, James J. 2005. Stochas- from genotypes to phenotypes. Nat. Rev. Genet., 6(June), ticity in gene expression: 451–464. Karig, David K, Siuti, Piro, Dar, Roy D, Retterer, Scott T, Doktycz, Mitchel J, and Simpson, Michael L. 2011. Model for biological communication in a nanofabricated cell-mimic driven by stochastic resonance. Nano communication networks, 2(1), 39–49. 19 Korn, Granino Arthur, and Korn, Theresa M. (eds). 1952. Electronic Analog Computers: D-c Analog Computers. McGraw-Hill. Latty, Tanya, and Beekman, Madeleine. 2013. Keeping track of changes: the performance ofa nt colonies in dynamic environments. Anim. Behav., 85(3), 637–643. McDonnell, Mark D., and Abbott, Derek. 2009. What is stochastic resonance? Definitions, misconceptions, debates, and its relevance to biology. PLoS Comput. Biol., 5(5), e1000348. Moss, Frank, Ward, Lawrence M, and Sannita, Walter G. 2004. Stochastic resonance and sensory information processing: a tutorial and review of application. Clinical neurophys- iology, 115(2), 267–281. Neiman, Alexander, Shulgin, Boris, Anishchenko, Vadim, Ebeling, Werner, Schimansky- Geier, Lutz, , and Freund, Jan. 1996. Dynamical entropies applied to stochastic resonance. Phys. Rev. Lett., 76(23), 4299–4302. Nicolis, C. 1982. Stochastic aspects of climatic transitions-response to a periodic forcing. Tellus, 14, 1–9. Nicolis, Stamatios C., and Deneubourg, Jean-Louis. 1999. Emerging patterns and food recruitment in ants: an analytical study. J. Theor. Biol., 198, 575–592. Penzias, A. A., and Wilson, R. W. 1965. A measurement of excess antenna temperature at 4080 Mc/s. Astrophys. J., 142, 419–421. Popat, R, Cornforth, DM, McNally, L, and Brown, SP. 2015. Collective sensing and collective responses in quorum-sensing bacteria. Journal of The Royal Society Interface, 12(103), 20140882. Reid, Chris R., Sumpter, David J. T., and Beekman, Madeleine. 2011. Optimisation in a natural system: ARgentine ants solve the Towers of Hanoi. J. Exp. Biol., 214(January 1,), 50–58. Roberts, Lawrence Gilman. 1962. Picture coding using pseudo-random noise. IRE Trans. Inf. Theory, 8(2), 145–154. Rolls, ET, and Deco, G. 2010. The noisy brain. Stochastic dynamics as a principle of brain function.(Oxford Univ. Press, UK, 2010). Sanchez, Alvaro, Choubey, Sandeep, and Kondev, Jane. 2013. Regulation of noise in gene expression. Annual review of biophysics, 42, 469–491. Schuchman, Leonard. 1964. Dither signals and their effect on quantization noise. IEEE Trans. Commun. Technol., 12(4), 162–165. Viney, Mark, and Reece, Sarah E. 2013. Adaptive noise. Proceedings of the Royal Society of London B: Biological Sciences, 280(1767). Wagner, Andreas. 2014. Arrival of the Fittest: Solving Evolution's Greatest Puzzle. Penguin. Waters, Christopher M, and Bassler, Bonnie L. 2005. Quorum sensing: cell-to-cell commu- nication in bacteria. Annu. Rev. Cell Dev. Biol., 21, 319–346. 20 Wilson, Edward O. 1962. Chemical communication among workers of the fire ant Solenopsis saevissima (Fr. Smith): 2. An information analysis of the odour trail. Anim. Behav., 10(1–2), 148–158. 21
1706.08013
1
1706
2017-06-24T23:53:36
Mapping The Ultrafast Flow Of Harvested Solar Energy In Living Photosynthetic Cells
[ "physics.bio-ph" ]
Photosynthesis transfers energy efficiently through a series of antenna complexes to the reaction center where charge separation occurs. Energy transfer in vivo is primarily monitored by measuring fluorescence signals from the small fraction of excitations that fail to result in charge separation. Here, we use two-dimensional electronic spectroscopy to follow the entire energy transfer process in a thriving culture of the purple bacteria, Rhodobacter sphaeroides. By removing contributions from scattered light, we extract the dynamics of energy transfer through the dense network of antenna complexes and into the reaction center. Simulations demonstrate that these dynamics constrain the membrane organization to be small pools of core antenna complexes that rapidly trap energy absorbed by surrounding peripheral antenna complexes. The rapid trapping and limited back transfer of these excitations lead to transfer efficiencies of 83% and a small functional light-harvesting unit.
physics.bio-ph
physics
Mapping The Ultrafast Flow Of Harvested Solar Energy In Living Photosynthetic Cells Peter D. Dahlberg1, Po-Chieh Ting2, Sara C. Massey2, Marco A. Allodi2, Elizabeth C. Martin3, C. Neil Hunter3, and Gregory S. Engel2* 1. Graduate Program in the Biophysical Sciences, Institute for Biophysical Dynamics, and the James Franck Institute, The University of Chicago, Chicago, IL 60637 2. Department of Chemistry, Institute for Biophysical Dynamics, and the James Franck Institute, The University of Chicago, Chicago, IL 60637 3. Department of Molecular Biology and Biotechnology, University of Sheffield, Firth Court, Western Bank, Sheffield S10 2TN, UK * Corresponding Author E-mail: [email protected] (773-834-0818) Abstract Photosynthesis transfers energy efficiently through a series of antenna complexes to the reaction center where charge separation occurs. Energy transfer in vivo is primarily monitored by measuring fluorescence signals from the small fraction of excitations that fail to result in charge separation. Here, we use two-dimensional electronic spectroscopy to follow the entire energy transfer process in a thriving culture of the purple bacteria, Rhodobacter sphaeroides. By removing contributions from scattered light, we extract the dynamics of energy transfer through the dense network of antenna complexes and into the reaction center. Simulations demonstrate that these dynamics constrain the membrane organization to be small pools of core antenna complexes that rapidly trap energy absorbed by surrounding peripheral antenna complexes. The rapid trapping and limited back transfer of these excitations lead to transfer efficiencies of 83% and a small functional light- harvesting unit. 1 Introduction Photosynthesis relies on ultrafast energy transfer to efficiently move energy from the site of absorption (photosynthetic antenna) to the site of charge separation (photosynthetic reaction center, RC) on a picosecond timescale.1,2 These processes are crucial for the organism's fitness and are tightly regulated and optimized for safe harvesting of solar energy.3-6 For the better part of a century, in vivo energy transfer has been monitored by measuring the small percentage of absorbed light emitted as fluorescence. Here, we reveal the fate of the majority of absorbed excitations in a living bacterial cell, using two- dimensional electronic spectroscopy (2DES) to follow the flow of energy in the model organism Rhodobacter sphaeroides. The light-harvesting machinery from Rba. sphaeroides is composed of two different types of antenna complexes, Fig. 1a.7,8 Light-harvesting complex 2, LH2, has two rings of bacteriochlorophyll a (Bchl a). One ring contains 9 weakly coupled Bchl a that absorb around 800 nm (Fig. 1b), known as the B800 band, and the other ring contains 18 strongly coupled Bchl a that absorb around 850 nm, known as the B850 band.7 Light- harvesting complex 1, LH1, dimerizes to form an "S" shaped array of 56 Bchl a that encircles two RCs. Dimerization is driven by the protein PufX that links two LH1s together at a slight angle that also drives membrane curvature.8 The 56 Bchl a in an LH1 dimer are strongly coupled and form a complex electronic structure with a dominant absorption feature at 875 nm, known as the B875 band. Studies on membrane fragments suggest that LH2 transfers energy to LH1 on a few picoseconds timescale and LH1 in turn transfers energy to the special pair of Bchl a in the RC, the site of charge separation.9-13 The transfer between LH1 and the special pair is energetically uphill, from 875 nm to 870 nm, and is the rate-limiting step in the energy transfer process, occurring within approximately 50 picoseconds.11,14 In this work, we overcome the longstanding obstacle of intensely scattered light15,16 and recover the native dynamics of energy transfer through the dense network of antenna complexes and into the reaction center. 2DES spectra correlate excitation energy along the ωτ-axis with stimulated emission (SE), ground state bleach (GSB), and excited state absorption (ESA) energies along the ωt-axis as a function of an ultrafast time delay, T, known as the waiting time.17,18 Here, we use the GRadient Assisted Photon Echo Spectrometer (GRAPES), which multiplexes one of the sampling dimensions.19 This multiplexing reduces signal acquisition time by several orders of magnitude and allows fine sampling of the waiting time domain. This fine sampling of the waiting time enables the removal of scattered light in Fourier domains inaccessible by conventional 2DES.16 The energy transfer dynamics observed in vivo in conjunction with simulations constrain the membrane organization to be small pools of core antenna complexes that rapidly trap energy absorbed by surrounding peripheral antenna complexes. These results agree with previous AFM,20 and ultrafast spectroscopy studies on isolated complexes and membrane fragments,9-14,21 as well as simulations and theoretical studies.22-25 Additionally, these results answer longstanding questions concerning the energy transfer dynamics and membrane organization truly present in living cells and are an essential demonstration that the isolation processes employed for decades do not significantly perturb the energy transfer process. Further, these measurements establish the powerful capabilities of this 2 variant of 2DES that can be broadly applied to more complex photosynthetic organisms such as plants and algae where numerous questions remain surrounding membrane architecture, dynamics, and regulatory processes. Results Annihilation dynamics in mutants containing only LH1 or LH2. Measuring energy transfer dynamics in well-connected systems of identical subunits, like those found in photosynthetic membranes, presents two key experimental challenges. First, much of the energy transfer occurs between isoenergetic complexes that are spectrally indistinguishable from one another and second, even at relatively low excitation concentrations exciton- exciton annihilation can dominate observed dynamics.2,26,27 Performing a power- dependence study of dynamics with excitation fluences of 17.6, 5.6, 3.6, and 2.2 μJ/cm2 per pulse allows us to identify annihilation processes by observing changes in dynamics between excitation fluences and then to exploit the annihilation dynamics to determine energy transfer times between the isoenergetic complexes. Fig. 1c shows the absorptive 2DES spectrum of cells containing only LH2 complexes28 at our highest excitation fluence of 17.6 μJ/cm2. The waiting time traces shown in Fig. 1d show clear power-dependent dynamics. From the molar extinction coefficient of LH2 and the excitation spectrum, we calculate that the lowest excitation fluences of 3.6 and 2.2 μJ/cm2 correspond to exciting ~1 in every 30 to 50 LH2 complexes respectively. The changing dynamics between these fluences is indicative of exciton-exciton annihilation and suggests a large number of connected LH2, in agreement with previous studies.27 The same power-dependent study is reproduced for cells containing only LH1 complexes28 (Supplementary Fig. 2) and shows annihilation only at the highest fluence corresponding to exciting ~1 in every 8 LH1 complexes. 3 Figure 1 Annihilation reveals a highly connected network of light‐harvesting complexes in vivo. a, Crystal structures of LH2 and RC‐LH1‐PufX (PDB 1KZU and PDB 4JC9 respectively) with the carotenoids and Bchl a phytyl tails removed for clarity. LH2 contains two bands of Bchl a, the B800 (blue) and the B850 (green). LH1 contains a single band of Bchl a, B875 (red) that transfers energy to the special pair of the RC (orange). b, Absorption spectra in a 200 μm path length of cells containing only LH2, cells containing only LH1, and wild type (WT) cells. The large offset from zero optical density is due to optical scattering. The 2DES excitation spectrum is shown in gray and is produced by super continuum generation in argon gas. The spectrum is broad enough to interrogate the entire energy transfer process from LH2LH1RC. c, Absorptive 2DES spectrum of LH2‐only cells taken with 17.6 μJ/cm2 at T = 1 ps. d, Waiting time traces acquired at different powers from the maximum of the GSB/SE feature. The traces are the average of 3 scans and the shaded background is the mean ± the standard deviation. The change in dynamics with power is indicative of exciton‐exciton annihilation. The dashed traces are the population of excited LH2 from a random walk simulation with a lifetime for energy transfer between LH2s of 2.7 ps, a domain size of 64 LH2, and a fluorescence lifetime of 250 ps. The annihilation dynamics are used to recover energy transfer times between isoenergetic complexes, which are normally unobtainable in 2DES studies because they yield no cross peaks or dynamic signals. Following the analysis of Barzda et al. on aggregates of an isolated antenna from plants2,26 the annihilation rate, γ0, is related to the transfer time, τhop, by (cid:2011)(cid:2868)(cid:2879)(cid:2869)(cid:3404)0.5(cid:1840)(cid:1858)(cid:3031)(cid:4666)(cid:1840)(cid:4667)(cid:2028)(cid:3035)(cid:3042)(cid:3043) (1) where N is the number of connected antenna complexes, in this case LH2, and fd(N) is the structure factor indicative of the packing arrangement. At large values of N, fd(N) depends weakly on N and is determined by the assumed arrangement of chromophores.26 Based on previous AFM studies,20 we assume an arrangement between a square lattice, fd = 0.8, and 4 a hexagonal lattice, fd = 0.6, and take fd(N) to be a constant value of 0.7. The 2DES signal can then be related to the annihilation rate via log(cid:4672) (cid:2869)(cid:3041)(cid:4666)(cid:3021)(cid:4667)(cid:3398) (cid:2869)(cid:3041)(cid:4666)(cid:2868)(cid:4667)(cid:4673)(cid:3404)(cid:4666)(cid:1856)(cid:3046)/2(cid:4667)log(cid:4666)T(cid:4667)(cid:3397)log(cid:4672)(cid:3082)(cid:3116)(cid:3031)(cid:3294)(cid:4673) (2) Where n(T) is the population of excitations at waiting time T, and ds is the fractal dimension, typically around 1.8, which accounts for diffusion being restricted in the membrane to the lattice of antenna complexes.29 between LH2s. Waiting time traces from the data collected at a fluence of 5.6 μJ/cm2 are Fig. 2 illustrates the experimental procedure for recovering the energy transfer time fit to the linear relationship of equation 2 and are shown in Fig. 2b. These fits are performed for each point within the dashed box in Fig. 2a. The recovered lifetime of 2.7 ± 0.1 ps is homogenous and the same for the ESA feature above the diagonal as well as the GSB/SE feature below the diagonal. Data collected at any of these excitation fluences could have been used to estimate the energy transfer time between LH2s, however the highest fluence used risks initially populating LH2s with multiple excitations that would give rise to annihilation not due to inter-complex transfer and the lowest fluence likely has little annihilation. Each of these effects could skew estimates of the energy transfer times. For completeness, Supplementary Fig. 3 shows the resulting estimates of transfer times recovered from all 4 fluences. The same analysis performed on LH2-only cells is performed in Supplementary Fig. 4 on the LH1-only cells acquired with 17.6 μJ/cm2, the only fluence where annihilation was observed, yielding a lifetime of 4.7 ± 0.2 ps for energy transfer from LH1 to LH1. Model of LH1- and LH2-only membranes. Using the recovered energy transfer times we constructed model photosynthetic membranes. The model membranes for LH1- and LH2- only cells were constructed based on previous AFM work that showed LH2 complexes in the absence of LH1s and RCs assemble with approximately hexagonal close packing.30 LH1-only mutants with PufX have been shown to form tubular membrane structures,31 but the LH1-only mutant used in this study lacks PufX, meaning that it will be composed primarily of monomers and will likely adopt a flat membrane structure. Initially, the model membranes are populated randomly with excitations based on the four laser fluences used. The excitations then undergo a random walk on the membrane by calculating the probability of hopping or fluorescing in a 10 fs time step. Annihilation was simulated when one complex received two or more excitations at a time by reducing the number of excitations to one. These trajectories yielded waiting time traces for the number of excited complexes that could be compared to the experimental data, Fig. 1d. The domain size and fluorescence lifetimes were varied, see Supplementary Figures 5-7. The optimal fluorescence lifetime for both LH1- and LH2-only membranes was found to be 250-300 ps. This value is in agreement with previous results32 as well as lifetimes recovered from our LH1-only mutant data taken at low fluences that showed no annihilation and fit to a lifetime of 264 ± 7 ps, see Fig 3e and Supplementary Fig. 8. While the fluorescence lifetime of the two complexes was similar the domain size was found to be significantly different. LH2-only cells were found to have a domain size of 64 complexes and LH1-only cells were found to have a domain size of 16 complexes, see Fig. 1d and Supplementary Fig. 2. 5 Figure 2 Annihilation dynamics constrain energy transfer times between isoenergetic antenna. a, Absorptive 2DES spectrum of LH2‐only cells at T = 1 ps collected at 5.6 μJ/cm2. The dashed box is analyzed further for the lifetime of energy transfer between LH2 complexes. b, Waiting time dynamics from the maximum GSB/SE feature presented following equation 2, where the 2DES intensity is n(T). The intercept of the linear relationship is used to retrieve the annihilation rate, γ0, which is used to recover the hopping time, τhop, given in equation 1. c, Color map of the recovered τhop. The contours and saturation of the color are given by the intensity of the 2DES signal at 1 ps. d, Histogram of the recovered lifetimes giving a mean of 2.7 ps for the lifetime of energy transfer between LH2s. Energy transfer dynamics in wild type Rba. sphaeroides. With the lifetimes for energy transfer between LH2s and between LH1s from the mutant cell studies, we can perform the same power-dependent study on wild type cells to recover energy transfer times from LH2 to LH1 and from LH1 to the RC. The study shows annihilation only at the highest excitation fluence with annihilation appearing in both LH2 and LH1 spectral features, Fig. 4. The lack of observed annihilation dynamics at lower fluences indicates rapid transfer to LH1 and trapping via the RC. The data collected at 5.6 μJ/cm2 is used to determine inter- complex transfer timescales to avoid complications due to annihilation. Fig. 3 shows waiting time traces from the peak intensity of the GSB/SE feature of LH2 and LH1 in the wild type cell data. Traces were fit to a bi-exponential function in the spectral regions corresponding to LH2 and LH1 and yielded energy transfer times of 4.8 ± 0.2 ps from LH2 6 to LH1 and 49 ± 3 ps from LH1 to the RC, in agreement with previous measurements on membrane fragments.9,10,21 Figure 3 Energy transfer and trapping in the complex network of light‐harvesting machinery. a, Waiting time traces taken from the peak locations of GSB/SE features of LH2 (blue) and LH1(orange) in the wild type Rba. sphaeroides. The 2DES spectra were normalized to 1 ps and fit to a biexponential function. The shaded background is the mean ± the standard deviation. b, Color map of the first lifetime from the biexponential fit. This lifetime in the region of LH2 corresponds to the energy transfer from LH2 to LH1. The saturation of the color as well as the gray contours is given by the intensity of the 2DES spectrum at T = 1 ps. c, same as b except for the second lifetime in the biexponential fit. The saturation and contours are coming from the T = 50 ps spectrum and the lifetime corresponds to LH1‐RC energy transfer. d, Histogram of the lifetimes within the dashed boxes of b. These lifetimes correspond to LH2 LH1 transfer times. e, Histogram of the lifetimes within the dashed box of c. These lifetimes correspond to LH1 RC transfer. The red histogram is from LH1‐only cells and represents the fluorescence lifetime of LH1, 264 ± 7 ps, and is consistent with the lifetime recovered from modeling. Combining experimental lifetimes, summarized in Supplementary Table 1, with modeling, we were able to constrain the membrane architecture and functional unit in WT Rba. sphaeroides. A model WT membrane was constructed in a similar manner to the LH1- and LH2-only membranes. The geometry of the model followed both EM and AFM work that has shown that the membrane forms vesicles ~ 50 nm in diameter and that LH1 exists in small domains of dimers.20,33 The ratio of LH1:LH2 was set to 1:1.8 to be consistent with the absorption spectrum, and the back transfer rate from LH1 to LH2 was set to ~120 ps to match the fluorescence spectrum in Supplementary Fig. 9. Trapping via the RC results in an oxidized special pair of bacteriochlorophyll responsible for charge separation. The oxidized special pair produces a GSB signal that overlaps the B875 GSB/SE signal. In the simulations, the signal strength of a bleached special pair was approximated to be 0.43 7 times the strength of GSB/SE of an LH1. This approximation is based on the molar extinction coefficients for the special pair and LH1 chromophores where the excitation was modeled as delocalized over an average of 2.5 B875 chromophores.34,35 The lifetime of an oxidized special pair is longer than the 200 ps trajectories simulated here, so an RC can only act as a trap for 1 excitation per trajectory. As in the LH1- and LH2-only models, the membrane is initially populated with excitations based on the four excitation fluences. The excitations then undergo a random walk. During each 10 fs time step an excitation can either transfer to another antenna complex, fluoresce, be trapped in an RC if the excitation is on an LH1 with an available RC, or annihilate with another excitation if they are on a single site thus reducing the number of excitations on the site to one. As in the LH1- and LH2-only models the wild type model yielded waiting time dynamics that can be compared to experimental results. The domain size of LH1 complexes in an embedded array of LH2 was varied, as was the overall number of connected complexes. A functional unit of 2 domains of 8 LH1 complexes embedded in 29 LH2 complexes was found to be most consistent with the 2DES data, see Fig. 4 and Supplementary Fig. 10. Discussion The efficiency of the model membrane was determined by exploring trajectories where the membrane was populated with a single excitation in an LH2. These trajectories represent the conditions expected during low-light growth. Fig. 4 a shows two of the five thousand simulated trajectories. These simulations yielded rapid transfer to the small domains of LH1. The limited back transfer rate meant that only 46% of excitations made one or more transfers from LH1 to LH2 and the majority of time before trapping or fluorescing was spent exploring the LH1/RC domains. These dynamics lead to efficient trapping of 83% of excitations via the RC. The limited back transfer from the small pools of LH1 also restricted the excitation from exploring a large membrane environment. Given an infinite membrane the excitation, on average, spent 61% of the time exploring one of the two nearest LH1 domains leading to a definition of the functional light-harvesting unit in low-light conditions being the domain shown in Fig. 4a. Organisms are inherently dynamic and adapt to their environment, but their responses to external conditions depend on them being alive. Photosynthetic organisms are no exception. The Rba. sphaeroides in this study were grown in low-light and semi-aerobic conditions and their membrane organization reflects this environment. Starved for photons, they expressed and assembled a highly efficient energy transfer pathway to funnel excitations into the RC, but this is not always the membrane organization present in Rba. sphaeroides. It is well established that the ratio of LH1 and LH2 change depending on light levels, thus altering the membrane architecture.36 Under aerobic conditions photoprotective mechanisms have been observed in LH1 that couple B875 to carotenoid dark states.5 In plants and cyanobacteria, photoprotective mechanisms, such as non- photochemical quenching and state transitions, alter ultrafast energy transfer pathways and membrane architecture in response to oxidative stress, high-light conditions, and dehydration. Many questions remain on exactly how the dynamics, coupling, and architecture change under different stresses. The work presented here is a direct measure of the flow of excited states in vivo, and reveals the energy transfer pathways and membrane architecture native to Rba. sphaeroides. These results confirm decades-old assumptions that measurements made in isolated complexes and membrane fragments 8 investigating ultrafast energy transfer and its regulation in living photosynthetic systems. resembled those of the functional unit in vivo and lay the ground work for future studies Figure 4 Modeling and experiment constrain membrane architecture. a, Model of the functional light‐ harvesting unit and two representative simulated trajectories for the initial conditions of a single excitation in LH2 with all RCs open. b, Absorptive 2DES spectrum of wild type cells taken with 17.6 μJ/cm2 at T = 1 ps. Waiting time traces taken from the spectral locations corresponding to c, LH1 d, LH2 and e, energy transfer from LH2 to LH1 acquired at different excitation powers. The traces are the average of 3 scans and the shaded background is ± the standard deviation. The dashed lines show the dynamics from the model membrane. The error between model and experiment at long times in c is likely due to GSB from the reaction center special pair and the deviation at short times in d is likely due to spectral overlap between the LH2‐LH1 cross peak and the LH2 diagonal feature. 9 Methods Optical apparatus: Two-dimensional electronic spectroscopy: The excitation spectrum was produced by focusing the output of a Ti:sapphire regenerative amplifier with a 5 kHz repetition rate, 2 W output, and a 30 fs pulse duration centered at 800 nm (Coherent Inc.) into argon gas at 4 psi above atmospheric pressure. The generated supercontinuum was then spectrally and temporally shaped using an SLM-based pulse shaper (Biophotonics Solutions). The 2DES spectra were collected using the GRadient Assisted Photon Echo Spectrometer, which encodes the coherence time delay spatially across the sample. The photon echo and local oscillator fields were spectrally resolved along the rephasing axis by a spectrometer (Andor Shamrock) and heterodyne detection was achieved using a high-speed CMOS camera (Phantom Miro). Waiting times were encoded using a motorized linear stage (Aerotech). Data acquisition and analysis: Raw interferograms from the camera initially have axes of coherence time and rephasing wavelength. Coherence times were sampled in 0.9 fs steps from -200 to 200 fs. Waiting time spectra were collected at 25 Hz in 0.1 fs steps for 50 fs surrounding the time points 1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 12, 14, 16, 18, 20, 25, 30, 35, 40, 50, 75, 100, and 200 ps. i.e. 500 frames were taken between 0.975-1.025 ps. These 500 frames were averaged to produce a single frame representing 1 ps. By scanning the waiting time delay over 50 fs, the optical scatter, which oscillates at the optical period in the waiting time domain, is suppressed during the averaging, making the technique relatively immune to artifacts from intensely scatter light. Cubic spline interpolation was then performed to convert the single average image to axes of coherence time and rephasing frequency. An FFT then produced the image in the coherence time and rephasing time domains. The data was apodized in the rephasing time domain with a Tukey Window of width 200 fs and alpha value 0.25 centered on the photon echo signal. Next, the data was apodized in the coherence time domain with a window of width 350 fs and zero-padded to achieve approximately square frequency pixels. Lastly, the signal was shifted to zero rephasing and coherence time and a 2DFFT was performed to produce unphased 2DES spectra in the coherence frequency rephasing frequency domain. The rephasing and nonrephasing signals were then phased separately to the 1 ps pump-probe spectra following the projection slice theorem (Supplementary Fig. 11). Because pump-probe spectroscopy does not benefit from the same scatter removal techniques as 2DES, pump-probe spectroscopy was performed on isolated membrane fragments. The real valued data from the addition of the rephasing and nonrephasing spectra gave the absorptive spectra used for analysis in this manuscript. All data were triplicated in rapid succession to produce a mean valued spectrum with errors generated by the standard deviation of the three measurements. During the acquisition of the three trials, the sample was illuminated while flowing for a total of 23 minutes and the complete experiment time was 1.5 hours per sample. All data analysis was performed using Matlab with custom software. Error analysis: After data analysis, which includes filtering in Fourier domains, each point in the 2DES spectra is no longer statistically independent. The number of independent 10 points was used to determine the standard error on the mean energy transfer and fluorescence lifetimes presented. Cell culture: Mutants containing only LH1 were obtained by genomic deletion of puc1BA and pufLMX. This mutant strain lacks LH2, RC, and PufX, leading to a preferentially monomeric state of LH1. Mutants containing only LH2 were obtained by the genomic deletion of the pufBALMX genes. Genomic deletions were performed following the protocol described in Mothersole et al.28 All cells were cultured semi-aerobically in the dark at 30 °C. Prior to 2DES analysis the cells were centrifuged (Eppendorf 5810 R 4000 rpm) and resuspended in 50/50 (v/v%) glycerol water. Analysis was performed in a 200 μm path length flow cell (Starna Cells Inc.) with a reservoir of 5 ml. Membrane fragments used only for pump-probe spectroscopy were obtained by disrupting the cells using a French press at 14000 psi. A slow spin was performed (12000 rpm JA 30.STI for 20 min) to remove the large cellular debris. The supernatant was diluted to an optical density of ~0.3 in the NIR region in the same 200 μm path length flow cell used for the 2DES analysis. Acknowledgments The authors would like to thank MRSEC (DMR 14- 20709), the DARPA QuBE program (N66001-10-1-4060), AFOSR (FA9550-14-1-0367), the DoD Vannevar Bush Fellowship (N00014-16-1-2513), the Alfred P. Sloan Fellowship, the Camille and Henry Dreyfus Foundation, and the Searle Foundation for supporting the work in this publication. S.C.M. acknowledges support from the Department of Defense (DoD) through the National Defense Science & Engineering Graduate Fellowship (NDSEG) Program. P.D.D. acknowledges support from the NSF-GRFP program, and the National Institute of Biomedical Imaging And Bioengineering of the National Institutes of Health under Award Number T32-EB009412. C.N.H and E.C.M. were supported by grant BB/M000265/1 from the Biotechnology and Biological Sciences Research Council (UK) and an Advanced Award from the European Research Council (338895). This research was also supported by the Photosynthetic Antenna Research Center (PARC), an Energy Frontier Research Center funded by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences under Award Number DE-SC 0001035. That grant provided partial support for C.N.H. References Blankenship, R. E. Molecular mechanisms of photosynthesis. (Blackwell Science, 2002). Amerongen, H. v., Valkunas, L. & Grondelle, R. v. Photosynthetic excitons. (World Scientific, 2000). Muller, P., Li, X. P. & Niyogi, K. K. Non‐photochemical quenching. A response to excess light energy. Plant Physiol. 125, 1558‐1566, (2001). Goss, R. & Lepetit, B. Biodiversity of NPQ. J. Plant Physiol. 172, 13‐32, (2015). 11 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 Šlouf, V. et al. Photoprotection in a purple phototrophic bacterium mediated by oxygen‐dependent alteration of carotenoid excited‐state properties. PNAS 109, 8570‐8575, (2012). Ruban, A. V., Johnson, M. P. & Duffy, C. D. P. The photoprotective molecular switch in the photosystem II antenna. Bba‐Bioenergetics 1817, 167‐181, (2012). Mcdermott, G. et al. Crystal‐Structure of an Integral Membrane Light‐ Harvesting Complex from Photosynthetic Bacteria. Nature 374, 517‐521, (1995). Qian, P. et al. Three‐Dimensional Structure of the Rhodobacter sphaeroides RC‐LH1‐PufX Complex: Dimerization and Quinone Channels Promoted by PufX. Biochemistry 52, 7575‐7585, (2013). Sundström, V., Pullerits, T. & van Grondelle, R. Photosynthetic Light‐ Harvesting:  Reconciling Dynamics and Structure of Purple Bacterial LH2 Reveals Function of Photosynthetic Unit. The Journal of Physical Chemistry B 103, 2327‐2346, (1999). Zhang, F. G., Gillbro, T., van Grondelle, R. & Sundstrom, V. Dynamics of Energy‐Transfer and Trapping in the Light‐Harvesting Antenna of Rhodopseudomonas‐viridis. Biophys. J. 61, 694‐703, (1992). Beekman, L. M. P. et al. Trapping Kinetics in Mutants of the Photosynthetic Purple Bacterium Rhodobacter‐Sphaeroides ‐ Influence of the Charge Separation Rate and Consequences for the Rate‐Limiting Step in the Light‐ Harvesting Process. Biochemistry 33, 3143‐3147, (1994). Hess, S. et al. Temporally and spectrally resolved subpicosecond energy transfer within the peripheral antenna complex (LH2) and from LH2 to the core antenna complex in photosynthetic purple bacteria. Proc. Natl. Acad. Sci. U. S. A. 92, 12333‐12337, (1995). Nagarajan, V. & Parson, W. W. Excitation energy transfer between the B850 and B875 antenna complexes of Rhodobacter sphaeroides. Biochemistry 36, 2300‐2306, (1997). Visscher, K. J., Bergstrom, H., Sundstrom, V., Hunter, C. N. & van Grondelle, R. Temperature‐Dependence of Energy‐Transfer from the Long Wavelength Antenna Bchl‐896 to the Reaction Center in Rhodospirillum‐rubrum, Rhodobacter‐sphaeroides (WT and M21 Mutant) from 77 to 177K Studied by Picosecond Absorption‐Spectroscopy. Photosynth. Res. 22, 211‐217, (1989). Dahlberg, P. D. et al. Communication: Coherences observed in vivo in photosynthetic bacteria using two‐dimensional electronic spectroscopy. JCP 143, 101101, (2015). Dahlberg, P. D., Fidler, A. F., Caram, J. R., Long, P. D. & Engel, G. S. Energy Transfer Observed in Live Cells Using Two‐Dimensional Electronic Spectroscopy. JPCL 4, 3636‐3640, (2013). Brixner, T. et al. Two‐dimensional spectroscopy of electronic couplings in photosynthesis. Nature 434, 625‐628, (2005). Hamm, P. & Zanni, M. T. Concepts and methods of 2D infrared spectroscopy. (Cambridge University Pres, 2011). 12 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 Harel, E., Fidler, A. F. & Engel, G. S. Real‐time mapping of electronic structure with single‐shot two‐dimensional electronic spectroscopy. PNAS 107, 16444‐ 16447, (2010). Bahatyrova, S. et al. The native architecture of a photosynthetic membrane. Nature 430, 1058‐1062, (2004). van Grondelle, R., Dekker, J. P., Gillbro, T. & Sundstrom, V. Energy transfer and trapping in photosynthesis. BBA ‐ Bioenergetics 1187, 1‐65, (1994). van Grondelle, R. & Novoderezhkin, V. I. Energy transfer in photosynthesis: experimental insights and quantitative models. Phys. Chem. Chem. Phys. 8, 793‐807, (2006). Ritz, T., Park, S. & Schulten, K. Kinetics of excitation migration and trapping in the photosynthetic unit of purple bacteria. J. Phys. Chem. B 105, 8259‐8267, (2001). Şener, M. et al. Photosynthetic Vesicle Architecture and Constraints on Efficient Energy Harvesting. Biophys. J. 99, 67‐75, (2010). Sener, M. et al. Forster Energy Transfer Theory as Reflected in the Structures of Photosynthetic Light‐Harvesting Systems. ChemPhysChem 12, 518‐531, (2011). Barzda, V. et al. Singlet‐singlet annihilation kinetics in aggregates and trimers of LHCII. Biophys. J. 80, 2409‐2421, (2001). Vos, M., van Dorssen, R. J., Amesz, J., van Grondelle, R. & Hunter, C. N. The Organization of the Photosynthetic Apparatus of Rhodobacter‐sphaeroides ‐ Studies of Antenna Mutants Using Singlet Singlet Quenching. BBA ‐ Bioenergetics 933, 132‐140, (1988). Mothersole, D. J. et al. PucC and LhaA direct efficient assembly of the light‐ harvesting complexes in Rhodobacter sphaeroides. Mol. Microbiol. 99, 307‐ 327, (2016). Bunde, A. & Havlin, S. Fractals and disordered systems. 2nd rev. and enlarged edn, (Springer, 1996). Olsen, J. D. et al. The organization of LH2 complexes in membranes from Rhodobacter sphaeroides. J. Biol. Chem. 283, 30772‐30779, (2008). Jungas, C., Ranck, J. L., Rigaud, J. L., Joliot, P. & Vermeglio, A. Supramolecular organization of the photosynthetic apparatus of Rhodobacter sphaeroides. EMBO J. 18, 534‐542, (1999). Hunter, C. N., Bergstrom, H., van Grondelle, R. & Sundstrom, V. Energy‐ transfer dynamics in three light‐harvesting mutants of Rhodobacter sphaeroides: a picosecond spectroscopy study. Biochemistry 29, 3203‐3207, (1990). Kiley, P. J., Varga, A. & Kaplan, S. Physiological and structural analysis of light‐ harvesting mutants of Rhodobacter sphaeroides. J. Bacteriol. 170, 1103‐ 1115, (1988). Monshouwer, R., Abrahamsson, M., van Mourik, F. & van Grondelle, R. Superradiance and Exciton Delocalization in Bacterial Photosynthetic Light‐ Harvesting Systems. The Journal of Physical Chemistry B 101, 7241‐7248, (1997). 13 35 36 Straley, S. C., Parson, W. W., Mauzerall, D. C. & Clayton, R. K. Pigment content and molar extinction coefficients of photochemical reaction centers from Rhodopseudomonas spheroides. Biochim. Biophys. Acta 305, 597‐609, (1973). Driscoll, B. et al. Energy transfer properties of Rhodobacter sphaeroides chromatophores during adaptation to Low light intensity. PCCP 16, 17133‐ 17141, (2014). 14 Supplementary Information: Mapping The Ultrafast Flow Of Harvested Solar Energy In Living Photosynthetic Cells Peter D. Dahlberg1, Po-Chieh Ting2, Sara C. Massey2, Marco A. Allodi2, Elizabeth C. Martin3, C. Neil Hunter3, and Gregory S. Engel2* 1. Graduate Program in the Biophysical Sciences, Institute for Biophysical Dynamics, and the James Franck Institute, The University of Chicago, Chicago, IL 60637 2. Department of Chemistry, Institute for Biophysical Dynamics, and the James Franck Institute, The University of Chicago, Chicago, IL 60637 3. Department of Molecular Biology and Biotechnology, University of Sheffield, Firth Court, Western Bank, Sheffield S10 2TN, UK * Corresponding Author E-mail: [email protected] (773-834-0818) 1 Supplementary Figure 1: Waiting time series of absorptive spectra from cells containing only LH2 (top), cells containing only LH1 (middle), and wild type cells (bottom) taken at 17.6 μJ/cm2. The spectra show significant overlap between LH2 and LH1 as well as clear energy transfer cross peaks between LH2 and LH1 in the wild type cells at T = 5 ps and T = 20 ps. 2 Supplementary Figure 2: (top) Absorptive 2DES spectrum of LH1-only cells taken with 17.6 μJ/cm2 at T = 1 ps. (bottom) Waiting time traces taken from the spectral location indicated acquired at different powers. The traces are the average of 3 scans and the shaded background is ± the standard deviation. The change in dynamics with power is indicative of exciton-exciton annihilation. The dashed lines show agreement to a model membrane with LH1 domain sizes of 16 complexes, a transfer time of 4.7 ps, and an excited state lifetime of 250 ps. 3 Supplementary Figure 3: Histogram of the energy transfer times between LH2s recovered from 2DES data of LH2 only cells at fluences of 17.6, 5.6, 3.6, and 2.6 μJ/cm2. 4 Supplementary Figure 4: a Absorptive 2DES spectrum of LH1-only cells at T = 1 ps collected at 17.6 μJ/cm2. The dashed box is analyzed further for the lifetime of energy transfer between LH1 complexes. b Waiting time dynamics from the maximum ground state bleach and stimulated emission feature presented following equation 2, where the 2DES intensity is n(T). The intercept of the linear relationship is used to retrieve the annihilation rate, γ0, which in turn is used to recover the hopping time, τhop, given in equation 1. c Color map of the recovered τhop. The contours and saturation of the color is given by the intensity of the 2DES signal at 1 ps. d Histogram of the lifetimes recovered in c giving a mean of 4.7 ps for the lifetime of energy transfer between LH1s. 5 Supplementary Figure 5: Comparison of dynamics from LH1-only cells and LH1-only model membranes with varying domain size and fluorescence lifetime. The best fit is found with 16 complexes and a fluorescence lifetime of 250 ps. The comparison of experiment and model with these parameters can be seen in Figure S2. 6 Supplementary Figure 6: Comparison of dynamics from LH2-only cells and LH2-only model membranes with varying domain size and fluorescence lifetime. The best fit is found with 64 complexes and a fluorescence lifetime of 300 ps. The comparison of experiment and model with these parameters can be seen in Figure 1d. 7 Supplementary Figure 7: Comparison of experimental dynamics to simulated dynamics recovered from model membranes with a 250 ps fluorescence lifetime and varying domain sizes in LH1-only (a and b) and LH2-only (c and d) cells. The clear deviation from the model is indicative of the tight constraint on domain sizes in both LH1- and LH2-only cells. Supplementary Figure 8: Color map of the second lifetime from the biexponential fit. This lifetime in the LH1-only cells corresponds to the lifetime of the excited state in LH1 in the absence of the RC trap and annihilation. The saturation of the color as well as the gray contours are given by the intensity of the 2DES spectrum of LH1-only cells at T = 50 ps. 8 Supplementary Figure 9: (left) Absorption spectra of WT, LH1-only, and LH2-only cells with the scatter removed by fitting a quadratic function to long wavelengths. The dashed line is the result of a weighted sum fit to the WT absorption spectrum revealing a ratio of 1.8 LH2 to LH1. (right) Fluorescence emission spectrum from WT cells excited at 800 nm. The relative ratio between fluorescence intensity of LH2 and LH1 was determined by fitting two Gaussian functions to the spectrum and revealed about 13% of fluorescence was from LH2. This was used to constrain the back transfer rate from LH1 to LH2 to be on the order of 120 ps. 9 Supplementary Figure 10: Comparison of experimental data from the stimulated emission and ground state bleach peak of LH1 to dynamics recovered from model WT membranes with LH1 domain sizes of a, 4 and b, 16. The deviation from the model is indicative of the tight constraint on LH1 domain sizes in WT membranes. 10 Supplementary Figure 11: Example of the phasing procedure showing the projection of the rephasing 2DES spectrum of LH2-only cells (solid) to the pump-probe spectrum (dashed) at T = 1 ps. Complex-Complex τhop (ps) 2.7 ± 0.1 LH2 LH2 LH2 LH1 4.8 ± 0.2 4.7 ± 0.2 LH1 LH1 49 ± 3 LH1 RC 2731 LH2 fluorescence 264 ± 7 LH1 fluorescence Supplementary Table 1: Energy transfer times between complexes recovered from annihilation studies in mutant Rba. sphaeroides and biexponential fits to low fluence scans. Because annihilation was present at all powers in the LH2 only cells, the fluorescence lifetime of LH2 from Hunter et al.1 was used for modeling. Hunter, C. N., Bergstrom, H., van Grondelle, R. & Sundstrom, V. Energy‐ transfer dynamics in three light‐harvesting mutants of Rhodobacter sphaeroides: a picosecond spectroscopy study. Biochemistry 29, 3203‐3207, doi:Doi 10.1021/Bi00465a008 (1990). 11 1
1606.08218
1
1606
2016-06-27T11:36:21
Strain-driven criticality underlies nonlinear mechanics of fibrous networks
[ "physics.bio-ph", "cond-mat.soft" ]
Networks with only central force interactions are floppy when their average connectivity is below an isostatic threshold. Although such networks are mechanically unstable, they can become rigid when strained. It was recently shown that the transition from floppy to rigid states as a function of simple shear strain is continuous, with hallmark signatures of criticality (Nat. Phys. 12, 584 (2016)). The nonlinear mechanical response of collagen networks was shown to be quantitatively described within the framework of such mechanical critical phenomenon. Here, we provide a more quantitative characterization of critical behavior in subisostatic networks. Using finite size scaling we demonstrate the divergence of strain fluctuations in the network at well-defined critical strain. We show that the characteristic strain corresponding to the onset of strain stiffening is distinct from but related to this critical strain in a way that depends on critical exponents. We confirm this prediction experimentally for collagen networks. Moreover, we find that the apparent critical exponents are largely independent of the spatial dimensionality. In a highly simplified computational model of network dynamics, we also observe critical slowing down in the vicinity of the critical strain. With subisostaticity as the only required condition, strain-driven criticality is expected to be a general feature of biologically relevant fibrous networks.
physics.bio-ph
physics
Strain-driven criticality underlies nonlinear mechanics of fibrous networks A. Sharma1,2, A. J. Licup1, R. Rens1, M. Vahabi1, K. A. Jansen3,4, G. H. Koenderink3, F. C. MacKintosh1,5 1Department of Physics and Astronomy, VU University, Amsterdam, The Netherlands 2Department of Physics, University of Fribourg, CH-1700 Fribourg, Switzerland 3FOM Institute AMOLF, Science Park 104, 1098 XG Amsterdam, The Netherlands 4Wellcome Trust Centre for Cell-Matrix Research, Faculty of Life Sciences, University of Manchester, Manchester M13 9PT, UK 5Departments of Chemical and Biomolecular Engineering, Chemistry and Physics, Rice University, Houston, TX 77005, USA (Dated: August 8, 2018) Networks with only central force interactions are floppy when their average connectivity is below an isostatic threshold. Although such networks are mechanically unstable, they can become rigid when strained. It was recently shown that the transition from floppy to rigid states as a function of simple shear strain is continuous, with hallmark signatures of criticality [1]. The nonlinear mechan- ical response of collagen networks was shown to be quantitatively described within the framework of such mechanical critical phenomenon. Here, we provide a more quantitative characterization of critical behavior in subisostatic networks. Using finite size scaling we demonstrate the divergence of strain fluctuations in the network at well-defined critical strain. We show that the characteristic strain corresponding to the onset of strain stiffening is distinct from but related to this critical strain in a way that depends on critical exponents. We confirm this prediction experimentally for collagen networks. Moreover, we find that the apparent critical exponents are largely independent of the spatial dimensionality. In a highly simplified computational model of network dynamics, we also observe critical slowing down in the vicinity of the critical strain. With subisostaticity as the only required condition, strain-driven criticality is expected to be a general feature of biologically relevant fibrous networks. PACS numbers: Disordered filamentous networks are ubiquitous in bi- ology. An important example of such networks is the extracellular matrix of in biological tissues which is pre- dominantly composed of a fibrous collagen scaffold [2]. One of the most important characteristics of such net- works is the coordination number or average connectiv- ity hzi. Networks with only central force interactions are unstable towards small deformation if the average con- nectivity is below the threshold value of hzi = 2d, where d is the dimensionality. This threshold is referred to as the isostatic point at which, as shown by Maxwell [3], the number of degrees of freedom are just balanced by the number of constraints, and the system is marginally sta- ble. As the average connectivity increases beyond 2d, the network undergoes a phase transition marked by a con- tinuous increase in the elasticity. Other examples of such transitions are the jamming transition [4 -- 7] in granular materials and rigidity percolation [8 -- 11] in disordered spring networks. Jamming exhibits signatures charac- teristic of both first- and second-order transitions, with discontinuous behavior of the bulk modulus and contin- uous variation of the shear modulus [6, 12, 13]. For networks of springs or fibers, the transition from floppy to rigid is a continuous phase transition, in both bulk and shear moduli, with critical signatures [6, 8, 14 -- 17]. In a biological context, the average connectivity is al- most always below the isostatic threshold. Filamentous networks typically fall in two categories, those in which network formation occurs via branching and those where crosslinking proteins connect two distinct filaments. The typical connectivity in such networks is between 3 and 4, with the former due to branching and the latter due to binary crosslinking. In fact these networks are well below both 2D and 3D isostatic thresholds [16, 18]. Such subisostatic networks can, however, become rigid as a re- sult of other mechanical constraints, such as fiber bend- ing [14, 16, 19, 20], internal stresses [21], thermal fluctu- ations [22], or when subjected to external strain [17, 23]. Except for the external strain, other applied fields sta- bilize the network even in the zero strain limit, i.e., the subisostatic network becomes stable to small deforma- tions. However, when the applied field is an external strain, the transition from floppy to rigid states occurs at a threshold strain which depends on the network struc- ture, nature of the applied deformation as well as the av- erage connectivity [17]. We recently showed that sheared subisostatic networks exhibit a line of second order tran- sitions at a strain threshold γc(z), for connectivities hzi well below the isostatic threshold [1]. Here we follow up on this intriguing finding of strain- driven criticality by performing a detailed study of the nonlinear mechanics under simple shear. As a hallmark signature of criticality we demonstrate the divergence of strain fluctuations in the thermodynamic limit using fi- nite size scaling. In Ref. [1] it was shown that the critical exponents appear to depend on the average connectiv- ity in the network. Here we present our findings on the evolution of critical exponents in more detail. As an- other probe of criticality, we examine whether subiso- static networks exhibit critical slowing down near the 2 104 ] a P [ x a m K 103 c 100 c [mg/mL] 101 (b) (a) Figure 1: (Color online) (a) Shear stiffness versus shear strain curves obtained from a phantom triangular lattice in 2D with hzi ≃ 3.4. Different curves are obtained by varying the reduced bending rigidity κ. The onset strain for stiffening γ0 is shown as the blue dash-dotted line. The red dashed line shows the stiffness when κ = 0. In absence of bending interactions, the stiffness remains zero for γ ≤ γc. The green dashed lines through the symbols show the predicted stiffness according to Eq. (9) with f = 0.8±0.05 and φ = 2.1±0.2. (b) Experimentally obtained stiffness versus strain curve for a 1mg/mL collagen network. Since K in simulations corresponds to K/c in experiments, the experimentally obtained stiffness is normalized to the concentration c. The dashed line through the experimental data is fit according to Eq. (9) with the parameters f = 0.8, φ = 2.3 obtained from the collapse of stiffness curves obtained from simulations as explained in Sec. II. The critical strain γc = 0.29, marked with a red cross, is obtained as the inflection point of the stiffness curve. The onset strain for stiffening γ0 is marked with a blue cross. The inset shows the experimentally measured Kmax versus concentration c for collagen networks, Kmax is the maximum nonlinear modulus before the network ruptures. At large strains, when network stiffness is governed by stretching, the network stiffness scales as Kmax ∼ c shown as the black line. critical strain. Using a simplified model of network dy- namics we find evidence for power-law dynamics near the critical point. The article is organized as follows. In Sec. I we de- scribe the computational model used in this study. We also describe the mapping of parameters used in simu- lations to the experimentally relevant control variables. In Sec. II we focus on the demonstration of strain-driven criticality in disordered networks. We show the critical scaling of the order parameter close to the critical point implying the continuous transition. In this section, we also analyse the stiffness versus strain curves for finite bending rigidities in terms of a crossover function. In Sec. II B, we investigate strain fluctuations at the critical point and demonstrate their divergence in the thermody- namic limit. In Sec. III, we derive an approximate equa- tion describing the shape of the stiffness versus strain curves. We show that the derived equation can accu- rately describe the mechanical response measured for re- constituted collagen networks. In Sec. IV, we obtain and experimentally validate scaling relation between the on- set strain for stiffening and the critical strain. In Sec. V, we show that under simple shear, the critical exponents vary with the average connectivity. In Sec. VI, we show that the dynamics of network relaxation are critically slowed down near the critical strain for simple shear. We discuss our findings together with an outlook in Sec. VII. I. THE MODEL We model lattice-based networks [24 -- 26] in 2D and 3D. Fibers are arranged on a triangular lattice (2D) or a face- centered cubic lattice (3D) of linear dimension W . In 2D, we randomly select two of the three fibers at each vertex on which we form a binary cross-link, i.e., enforcing lo- cal 4-fold connectivity of the network in which the third fiber does not interact with the other two [24]. Similarly, in 3D, where there are 6 fibers crossing at a point, we randomly connect three separate pairs of fibers at each vertex with binary cross-links to enforce local 4-fold con- nectivity [25]. In both 2D and 3D, the average connec- tivity is further reduced below 4 by random dilution of bonds with a probability (1 − p), where p is the probabil- ity that a bond exists. The resulting connectivity after dilution can be estimated as hzi ≃ 4p. All networks, by construction, are subisostatic and floppy in the absence of bending interactions [16]. The filaments are character- ized by both a stretching modulus, µ, and bending rigid- ity, κ. These define a dimensionless rigidity κ = κ/µl2, where l is the lattice spacing (mesh size) in lattice-based (Mikado) networks. In lattice-based networks we take l = l0 where l0 is the lattice constant. The networks are subjected to an affine simple shear strain γ and sub- sequently allowed to relax by minimization of the total elastic energy. The total elastic energy per unit volume, H, is calculated using a discrete form of the extensible 2 ds# , (1) κ 2 f(cid:12)(cid:12)(cid:12)(cid:12) dt ds(cid:12)(cid:12)(cid:12)(cid:12) wormlike chain Hamiltonian [27] H = 1 W dXf " µ ds(cid:19)2 2 f(cid:18) dl ds + where the term in the square brackets represents the en- ergy stored in a single fiber and the sum is performed over all the fibres in the networks. There are other choices of modelling an individual fiber such as a truss, Euler- Bernoulli or Timoshenko beam [28, 29]. The Hamiltonian in Eq. (1) captures the semiflexible nature of biopolymers with finite resistance to both tension and bending. De- tails about discretization of the Hamiltonian in Eq. (1) are described elsewhere [18]. The stress and modulus are obtained by taking first and second derivatives of the en- ergy density with respect to the applied deformation, re- spectively. The elastic energy involves a summation over all fibres in the network and is a function of the strain γ and the reduced bending rigidity κ. Since the modulus K involves the energy per unit volume, K is naturally proportional to the line density ρ defined as the total length of the fibers per unit volume [20, 27, 30 -- 32]. The modulus can therefore be expressed as K = µρK (γ, κ) , (2) where K is a function of the reduced bending rigidity and the applied deformation. From the computational perspective, the most relevant quantity is the function K (γ, κ). Consistent with our previous studies [1, 18, 33], we report the modulus (stress) in units of µρ. The line density ρ is specific to the chosen network architecture, i.e., the network geometry. In lattice-based networks, and ρ3D = 12p ρd = ρd/ld−1 [18]. For √2 Mikado networks, because of the polydispersity of lc it is more convenient to express the line density in terms of fiber length L such that ρM = ρM/L, where ρM = nf L2 and nf is the number of rods per unit area [34]. with ρ2D = 6p √3 c A. Relationship between model and experimental parameters In order to map our model onto experimental parame- ters, we make three basic assumptions: (1) the filaments are athermal, (2) the filaments behave as rods with a ho- mogenous elasticity, and (3) the network connectivity re- mains below the isostatic threshold throughout the range of polymerization conditions. Collagen networks, in gen- eral, satisfy these assumptions. Collagen fibers are rather thick and thermal fluctuations are therefore unlikely to play a significant role. As for the network connectiv- ity, we have experimentally verified for the concentration range 0.5 − 4 mg/mL and at two temperatures T = 30◦ and 37◦C that it remains below the isostatic threshold (see Fig. 2(b)). The most relevant experimental control variable is the total protein concentration c. For a given thickness of 3 fibers, the volume fraction ϕ of a network scales linearly with c and using the above assumptions can be simply related to the reduced bending rigidity κ as ϕ ∼ κ [1, 33, 35, 36]. It follows that K/ϕ (or K/c) in experiments can be directly compared with K (γ, κ) in simulations. Our theoretical results depend on the bending rigidity through the parameter κ which, as shown above, scales linearly with the protein concentration in experiments. This has an important consequence for the experimental rheology results; the magnitude of modulus and stress as well as the functional dependence of the stiffness on the applied deformation are insensitive to the fibril thickness for a given concentration. This can be understood as fol- lows. For a given total protein concentration, κ = κ/µl2 is insensitive to changes in fibril thickness since κ ∝ a4, µ ∝ a2 and l ∝ a. The structure of collagen networks, in- cluding fibril thickness, mesh size, homogeneity, and pre- sumably connectivity, depends in detail on concentration and polymerization conditions in nontrivial ways [37, 38]. However, under the basic assumptions mentioned above, κ remains a constant. II. STRAIN DRIVEN CRITICALITY In Fig. 1(a), we show the network stiffness K as a func- tion of the applied strain γ for different values of κ. In the inset to Fig. 1(a), we show that the linear modulus scales linearly with κ. The scaling K ∼ κ in the lin- ear regime has been reported in several computational studies [1, 14, 16, 19, 20, 33, 39]. That the computa- tional model is suitable for studying athermal networks such as collagen is based on the following observations. (1) The computationally obtained modulus is in units of µρ (Eq. (2)) implying that G0 ≡ K(γ = 0) ∼ ρκ ∼ ρ2 consistent with experimental data sets on reconstituted networks of collagen type I [1, 33, 40, 41]. (2) As can be seen in Fig. 1(a) the onset of nonlinearity occurs at a strain γ0 which appears to be independent of κ. Ex- perimentally this corresponds to γ0 being independent of the total protein concentration which is indeed what has been observed in several studies [1, 33, 40, 41]. (3) For large strains, K is independent of κ, implying that K ∼ ρ, which is expected in the regime where strains are large enough to cause stretching of fibers. In order to verify if experiments indeed show the linear scaling of K with concentration we consider how Kmax varies with the con- centration c. Kmax is the nonlinear modulus of a network before undergoing failure due to the applied stress. As can be seen in the inset of Fig. 1(b), the experimentally measured Kmax scales linearly with the concentration im- plying that for large strains the nonlinear stiffness scales as K ∼ ρ. When fiber bending costs no energy, i.e., κ = 0, the stiffness K remains zero for strains γ ≤ γc. Above γc, K increases continuously from zero for κ = 0. The criti- cal strain γc is determined by the network architecture, in particular its average connectivity [17]. At the iso- γc (z) Floppy i n a r t s l a c i t i r C 0.5 0.4 0.3 0.2 0.1 c γ Rigid Connectivity zc T =37 ◦ C T =30 ◦ C (a) 2D 3D (b) 0.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 <z > Figure 2: (Color online) (a) Schematic diagram of the phase behavior of disordered fibrous networks. The curve γc(z) is the boundary between floppy and rigid states. (b) γc versus average connectivity for phantom triangular networks in 2D and FCC lattice based 3D networks. The critical strain de- creases with increasing connectivity and approaches zero at the isostatic threshold hzi = 2d. The shaded region spans connectivities in the range 3.0 − 3.6. The two open symbols correspond to γc of collagen networks prepared at 4mg/mL for two different polymerization temperatures. The symbols show average of 3 samples and error bars represent standard deviations. Per sample at least 100 junctions were measured to determine hzi. static threshold of hzi = 2d, a central force network is marginally stable with γc = 0. In Fig. 2(a), we show a schematic of the phase diagram in γ-z plane. For a given average connectivity below the isostatic threshold, in- creasing the deformation beyond γc causes a phase tran- sition from floppy to rigid phase. The continuous curve γc(z) marks the boundary between the floppy and rigid states of subisostatic networks. In Fig. 2(b) we show the computationally obtained γc versus the average con- nectivity in the network when subjected to simple shear deformation. The critical strains for both 2D and 3D 4 Figure 3: Shear stiffness K versus ∆γ = γ − γc for different κ obtained from simulations on a phantom triangular network in 2D with hzi ≃ 3.4. In the limit of κ → 0, the stiffness K increases as a power-law in ∆γ with the critical exponent f , which for the given network is ≃ 0.77. networks are in quantitative agreement as long as the connectivity is suffciently below the 2D isostatic point hzi = 4. The shaded region in Fig. 2(b) spans the con- nectivities relevant for collagen networks. Also shown are two values of γc of 4mg/mL collagen networks for two temperatures. The critical strain is in quantitative agreement with the model for T = 30◦C. The appar- ent disagreement for T = 37◦C is probably due to the uncertainty associated with determination of hzi in the experiments. It is possible that due to finite resolution in experiments, the connectivity at some of the nodes is measured as 4 due to overlapping collagen fibers. This would lead to an overestimation of hzi and can thus ac- count for the disagreement between theory and experi- ments at T = 37◦C. In Fig. 3, the network stiffness is shown for several val- ues of κ in the vicinity of the critical strain. The contin- uous nature of the transition from floppy to rigid states is evident in the critical scaling of the network stiffness K ∼ ∆γf where ∆γ = γ − γc ≥ 0 and f is a criti- cal exponent. As shown in Fig. 3, the power-law scaling of stiffness is apparent only in the limit of κ → 0. Ex- tracting f as the limiting slope of K vs. ∆γ provides an independent method of obtaining this critical exponent. The same exponent can be obtained by scaling analysis as described in Sec. II A. Power-law scaling of the order parameter, K (or K) in our case, is a hallmark signature of critical phenomena. In fact, the strain-driven phase transition is strictly defined only for κ = 0 at which the interactions within the network are purely central force interactions. Upon addition of a field such as fiber bend- ing, the network becomes stable for γ < γc with the stiffness K ∝ κ. In absence of bending interactions, the phase behavior characterized by the continuous transition of the order 5 where G± is a scaling function with the positive and neg- ative branches corresponding to ∆γ > 0 and ∆γ < 0, respectively. This scaling is analogous to that for the conductivity of random resistor networks and fiber net- works as a function of connectivity [16, 42]. In Fig. 4, we test this by plotting K∆γ−f vs. κ∆γ−φ, according to Eq. (3). For x ≪ 1, G+(x) is approximately con- stant and G−(x) ∝ x. That G+(x) is approximately constant for x ≪ 1 captures the critical scaling of K as K ∼ ∆γf . The scaling G−(x) ∝ x captures the bend-dominated linear modulus where the linear modu- lus scales as K ∼ κ. Since K must be finite at ∆γ = 0, we also expect K ∼ κf /φµ1−f /φ, consistent with Eq. (3). We show in Fig. 4 the data obtained from phantom triangu- lar networks in 2D (same as in Fig. 1(a)) and FCC-based 3D lattices collapsed according to Eq. (3). Interestingly, the data collapse with the same exponents f ≃ 0.8 and φ ≃ 2.1. The average connectivity for the two differ- ent networks is chosen to be ≃ 3.4. Data from Mikado networks with the same average connectivity hzi as in lattice-based networks can be collapsed with the same critical exponents [1]. In fact, as we show in Sec. V, the exponents appear to be independent of the spatial dimensionality and are primarily determined by the av- erage connectivity. Mapping protein concentration to κ as described in Sec. I A allows us to obtain an analogous scaling relation applicable to experimental data. Since computationally one obtains K one must create the analogous quantity in experiments by scaling the measured modulus with concentration, i.e., K/c. On substituting c for κ and K/c for K in Eq. (3), we obtain the scaling function to collapse the experimental data as shown by us in Ref. [1]. The scaling function G±, with f and φ as input param- eters, describes the stiffening curves over the entire elastic regime for any concentration (or κ in simulations). One can obtain an analytical G± (approximately) exploiting the analogy of nonlinear mechanics to ferromagnetism as we show in Sec. III. Figure 4: (Color online) (a) Collapse of shear stiffness versus shear strain curves of Fig. 1(a) according to Eq. (3). Simula- tion data from 3D network with same connectivity as in 2D of hzi ≃ 3.4 collapse with the same critical exponents f = 0.8 and φ = 2.1. parameter K is reminiscent of the ferromagnetic phase transition. Magnetic materials are characterized by a Curie temperature Tc such that for T above Tc, the ma- terial is paramagnetic. On lowering the temperature T below Tc, there is spontaneous magnetization M of the material which increases continuously from zero as M ∝ ∆T β where ∆T = T − Tc < 0 and β is the criti- cal exponent. Above the Curie temperature, the param- agnetic phase is characterized by a zero magnetization. However, in presence of a finite magnetic field H, there is a net magnetization in the paramagnetic phase with M ∝ H. It is an intriguing analogy that by mapping κ to external field H and γ to the temperature T , one can study the transition from floppy to rigid states the same way as in a ferromagnet as further elaborated in Sec. III. A. Crossover for finite κ B. Divergent fluctuations The power law scaling of K with ∆γ is a hallmark signature of criticality and is strictly observed only when κ = 0. It is obvious that in this regime, the modulus is entirely governed by stretching of fibers. For any finite κ, a subisostatic network is stable for ∆γ < 0. In fact, for sufficiently small κ, the linear modulus of a subisostatic network is bending governed leading to K ∼ κ for γ < γc [1, 14, 16, 19, 20, 33]. Analogous to ferromagnetism, in presence of finite auxiliary field κ, the network undergoes a strain driven crossover from the bend dominated regime ∆γ < 0 to the stretch dominated regime ∆γ > 0. These two regimes can be summarized by the scaling form K ∝ ∆γf G±(cid:18) κ ∆γφ(cid:19) , (3) In a thermal critical phenomenon, there are diver- gent fluctuations in the order parameter at the critical point. In the athermal network under consideration in this study, there are no divergent fluctuations in the macroscopic K. One can, however, measure fluctuations by considering the deviation of the strain field within the network from the expected affine field [32, 43]. Under affine deformation, filaments are either stretched or com- pressed. Deviations from the affine deformation induce bending on filaments which can be considered as a mea- sure of fluctuations. These fluctuations are suppressed by a finite field such as κ. In Fig. 5(a), we plot the bending angle θijk averaged over the entire network for different values of κ. The triplet {i, j, k} corresponds to three consecutive crosslinks labeled as i, j, and k and the 6 κ =10−3 κ =10−4 κ =10−5 κ =10−6 κ =10−7 (d) (b) 10-1 γ κ =10−3 κ =10−4 κ =10−5 κ =10−6 κ =10−7 (c) 101 θ Γ 100 10-1 2.0 1.5 ν / λ − W θ Γ 1.0 0.5 γ (a) 10-1 W= 40 W= 60 W= 80 W= 100 W= 150 W= 200 W= 250 10-1 γ 0.0 −0.4 −0.3 −0.2 −0.1 0.0 0.1 0.2 0.3 0.4 W1/ν ∆γ 10-1 (cid:0) 2 θ (cid:1) 10-2 101 θ Γ 100 10-1 Figure 5: (Color online) Divergent fluctuations at the critical strain. (a) Average bending angle hθ2i obtained from simulations on a phantom triangular network in 2D with hzi ≃ 3.4 for different values of κ (see legend). The network size is W 2 = 2502. The thick black line indicates the expected small-strain γ2 scaling. hθ2i increases monotonically with γ. The shaded region is approximately the range γc − γ0. In this range, the rate of increase of hθ2i is strongly dependent on κ. (b) Γθ(γ) obtained as the derivative of data in (a) with respect to γ. In the limit of κ → 0, Γθ diverges at γ = γc. (c) Γθ versus γ for different system sizes (see legend). The bending rigidity is κ = 10−7. (d) Collapse of data in (c) according to Eq. (5) with λ = 0.6 ± 0.1 and ν = 2.0 ± 0.1. average implies summing over all the triplets in the net- work. As can be seen in Fig. 5(a), the average bending angle increases with the applied deformation. For small strains, the increase is quadratic in γ as expected in the linear regime. At large strains, the average bending an- gle increases very slowly with the applied deformation. In the intermediate strain range, shown as the shaded region, the rate of increase of average bending angle de- 7 Figure 6: (Color online) Non-affine displacements in a 2D phantom triangular network with hzi ≃ 3.4 and κ = 10−6 are shown as the network is deformed through the critical strain γc. The arrows indicate the deviation of a node from the imposed deformation. The magnitude of the vectorial displacements is largest at the critical strain. The color bar on the right indicates the elastic energy in bending (green) or stretching (red) form. pends strongly on κ. We define Γθ as the rate of change of the average bending angle with the applied strain. Γθ(γ) = ∂hθ2 ijki ∂γ . (4) In Fig. 5(b), we plot Γθ as a function of γ for different values of κ. These results are obtained from simulations on a phantom triangular network in 2D with hzi ≃ 3.4. The maximum of Γθ shifts to the left in γ with decreasing bending rigidity. In the limit of κ = 0, the peak height is maximum for a given network size and it is located at the critical strain γ = γc. The quantity Γθ is expected to diverge in the thermo- dynamic limit at γ = γc for κ = 0. In Fig. 5(c), we show Γθ for different system sizes W . These curves are obtained for a fixed small κ = 10−7. If Γθ diverges as γ − γc−λ in the thermodynamic limit W → ∞, then the following scaling relation must capture the scaling behavior of Γθ for finite W : Γθ ∝ W λ/ν H(W 1/ν∆γ), (5) where ν is the exponent associated with the divergence of correlation length [1], ∆γ = γ−γc is the distance from the critical strain and H(x) is a scaling function. We show in Fig. 5(d), the collapse of data in Fig. 5(c) according to Eq. (5) with the exponents λ = 0.6±0.1 and ν = 2.0±0.1. With these exponents, the peak height of Γθ is expected to scale as W λ/ν ∼ W 0.3. It follows that due to the weak system size dependence, a clear demonstration of W λ/ν scaling of the peak height requires much larger system sizes than those studied in this work. Nevertheless, the collapse in Fig. 5(d) provides convincing evidence for Γθ as an appropriate measure of fluctuations in fibrous net- works. Another measure of fluctuations is the differential non- affinity which measures the strain fluctuations within the network. Given the displacement field u and the affine displacement field uA of the network, the non-affine fluc- tuations can be quantified as [17] δΓ(γ) = hkδuNAk2i l2dγ2 , (6) where δΓ(γ) is referred to as differential non-affinity, δuNA = u − uA is the differential non-affine displacement of a crosslink to an imposed strain dγ, l is the typical network mesh size and the angular brackets represent a network average. In Ref. [1], we showed that δΓ(γ) ex- hibits a peak at γ = γc, the height of which increases with decreasing κ. In Fig. 6, we show the differential non-affine displacements δuNA superimposed on network nodes in the neighborhood of γc. The magnitude of non- affine displacements is largest at the critical strain. It follows that the network is at its most susceptible me- chanical state at γ = γc requiring large scale internal rearrangements in response to an infinitesmal external deformation. The nature of deformation within the net- work changes dramatically when the applied deformation increases through γ = γc. Whereas the network deforms primarily through bending modes for γ ≤ γc, stretching becomes the dominant deformation mode for γ > γc. Finite-size scaling analysis of the order parameter K reveals underlying divergence of the correlation length as shown in Ref. [1]. The diverging correlation length, to- gether with divergent fluctuations and the continuously evolving order parameter constitute evidence in favor of a second-order type strain-driven phase transition in dis- ordered networks. III. EQUATION FOR THE CROSSOVER FUNCTION The scaling ansatz and function G±(x) in Eq. (3) can account well for the nonlinear mechanics of our model networks for any κ and γ. We can obtain an analyti- cal approximation for G±(x) in a way analogous to the approach for ferromagnetism [44, 45]. In a way simi- lar to the equation of state relating magnetic field H to magnetization M , we postulate the following mean-field equation of state for bending stiffness κ and as a series in the shear modulus K [46]: κ ∼ bK + cK2, (7) where b ∼ ∆γ for a transition controlled by strain. Here, in contrast with the order parameter M for ferromag- netism, symmetry does not forbid a quadratic term in this equation of state [46]. After a minor change in nor- malization, this can be rewritten as κ ∆γ2 ∼ K ∆γ(cid:18)∓1 + K ∆γ(cid:19) , (8) where the upper '−' refers to γ > γc and the lower '+' refers to γ < γc. This yields K ∼ ∆γ for small ∆γ > 0 and κ = 0, while K ∼ κ for ∆γ < 0 and small κ > 0. As shown above, our results deviate from the mean-field behavior, K ∼ ∆γf , where f = 1. We find f ≃ 0.8. As is done for ferromagnetism, the equation of state above can be written in a form that can account for non- mean-field exponents, while remaining non-singular ex- cept at the critical point (∆γ = κ = 0). We introduce potentially non-integer exponents f and φ, where κ ∆γφ ∼ K ∆γf (cid:18)∓1 + (φ−f ) K1/f ∆γ(cid:19) . (9) For ∆γ = 0, this scaling relation corresponds to K ∼ κf /φ at the critical point. Again, the mean-field values of the exponents are f = 1 and φ = 2. Equation (9) can be used to calculate K for any γ. The input parameters are κ, f , φ and γc. The critical strain γc can be independently determined from a network with only central-force interactions. The critical exponents are obtained from the data collapse using Eq. (3). In Fig. 1(a), we use Eq. (9) to obtain K as a function of γ for different κ. The stiffening curves calculated using Eq (9) are shown together with the numerically obtained curves. Clearly, Eq. (9) can accurately predict the non- linear stiffening curves. Equation (9) can accurately capture the experimen- tally obtained stiffening curves of collagen networks [1]. However, the fitting procedure, when applied to experi- ments needs to be slightly modified. The fitting to exper- imental data is done in the following way. We first focus on the linear regime. In the linear regime, we know from simulations that the modulus (in units of ρµ) scales lin- early with κ which itself scales as κ ∼ ρ giving rise to a c2 8 c Figure 7: (Color online) The onset strain for stiffening scales as γ0 ∼ γ(φ−f ) . The critical exponents are φ = 2.1 and f = 0.8. The experimental data are taken from collagen networks prepared at temperatures, T = 30◦C (◦) and 37◦C ((cid:3)). This scaling is a direct consequence of the measured c2 scaling of the shear stress at γ0 as shown in the inset. (or ρ2) dependence of the linear modulus where c is the protein concentration. However, as shown in the inset of Fig. 1(b), the linear modulus obtained experimentally from reconstituted collagen networks exhibits K ∼ c2+δ scaling. It is plausible that the deviation from the c2 scal- ing is simply a consequence of experimental uncertainties. However, as shown in Ref. [1], the deviation from c2 scal- ing is probably due to the weak dependence of γc on the concentration of collagen in experiments. In this sec- tion, we simply rescale the experimental K by c1+δ such that rescaled modulus scales as K/c1+δ ∼ c ∼ κ. Next, we obtain the individual critical strains, γc, for each of the concentrations as the inflection point of the log K vs. log γ curve. We then consider the experimental data (rescaled by c1+δ) for each concentration along with its γc and fit the entire curve to Eq. (9) with κ as the only free parameter. Here we show the result of the fitting for a 1mg/mL collagen network in Fig. 1(b) superimposed on the experimental data. We have reported the full set of experimental curves over a wide range of concentrations of collagen along with the fitting in Ref. [1]. IV. RELATION BETWEEN γ0 AND γc In a recent study, we showed that the onset of stiffening strain γ0 is practically independent of the concentration of collagen [33]. The invariance of the geometrical struc- ture of the network with concentration, in particular of the average connectivity in the network, was suggested as the underlying reason for the independence. The same argument leads to the conclusion that γc is independent of the concentration and should be determined entirely 9 V. CRITICAL EXPONENTS AND CONNECTIVITY Strikingly, the critical exponents obtained by collaps- ing both simulation data of 2D and 3D fibrous networks and experimental data for collagen networks are identi- cal [1]as long as the average network connectivity is the same. The exponents are apparently independent of the spatial dimensionality. This is in contrast to both ther- mal and athermal critical phenomena where the critical exponents depend on the spatial dimensionality [45, 47]. In fact, the critical exponents evolve with the average connectivity in the network. In Fig. 8, we show the nonlinear stiffness data collapsed according to Eq (3) for 2D triangular lattice-based networks prepared at differ- ent connectivities. The inset of Fig. 8 shows a plot of f and f /φ versus the average connectivity for both 2D and 3D lattice-based networks. It is clear that f increases with the average connectivity in the network whereas φ remains practically constant. The evolution of critical ex- ponents with the connectivity has been also observed in branched networks modeled as diluted honeycomb struc- tures [48]. The continuous variation of critical exponents is sim- ilar to the behavior of Ashkin-Teller and 8-vertex mod- els, which exhibit continuously varying critical exponents along a critical line [49 -- 51]. Such a variation in the crit- ical exponents has been experimentally observed in cer- tain quantum phase transitions [52, 53]. In Ref. [48], we presented a hypothesis that the apparent variation of the critical exponents could correspond to a crossover be- tween critical exponents in the pure and disordered lim- its where the pure limit corresponds to an undiluted and undistorted perfect lattice based network. At present it remains unclear whether the variation can be attributed to a crossover behavior. However, based on previous sim- ulations [17] an interesting experimental verification of varying exponents could be to isotropically compress a subisostatic random network, since this would reduce the number of constraints while leaving the connectivity the same. VI. CRITICAL SLOWING DOWN One of the hallmark signatures of a critical phe- nomenon is extremely slow dynamics at the critical point [45]. The dynamics are characterized by a divergent relaxation time scale. In a disordered fibrous network, we investigate the critical slowing down by applying an affine deformation to the network such that the strain equals the critical value. We only take central-force interactions into account by setting κ = 0. We then let the net- work relax the elastic energy by performing overdamped Molecular Dynamics simulations. We do not take hydro- dynamics into account. We also ignore the asymmetric nature of drag acting on each filament. We rather assume that the drag forces acting on the network due to the sur- Figure 8: (Color online) Shear stiffness versus shear strain curves collapsed according to the Eq. (3) for phantom trian- gular networks in 2D prepared at different connectivities (see legend). The red and the blue data sets have been shifted by a decade up and down, respectively, for better visualization. The exponent f changes significantly with hzi. With φ show- ing practically no dependence on the connectivity, the ratio f /φ increases with the connectivity as shown in the inset. by the geometry of the network. It is therefore expected that a general relation exists between γ0 and γc. An expression for γ0, based on geometrical arguments has been derived in Ref. [18]. We can obtain an expres- sion for γc in terms of γ0 and critical exponents in the following way. Using Eq. (3), the linear modulus G0 can be written as G0 ≡ K(γ = 0) ∼ c2γf−φ It follows that the stress at the onset of stiffening should scale as σ0 = G0γ0 ∼ c2γf−φ γ0. The experimentally obtained σ0 versus concentration is shown in the inset of Fig. 7. The data are taken from collagen networks prepared at tem- peratures, T = 30◦C and 37◦C. As can be seen in Fig. 7, σ0 scales quadratically with the concentration implying that . c c γ0 ∼ γ(φ−f ) c . (10) This scaling relation accurately describes the relation be- tween γ0 and γc as shown in Fig. 7 with φ = 2.1 and f = 0.8. However, unlike γ0, which can be determined analytically, determination of γc from Eq. (10) requires the knowledge of the critical exponents which, at present, are only obtained from scaling analysis of stiffening data. It is important to note that the above arguments are valid only when the average connectivity in the network depends weakly on the concentration. This requirement is based on the observation, as shown in the next section, that the critical exponents evolve with the average con- nectivity in the network. Using a unique set of values for φ and f in Eq. (10) requires that these two exponents are practically constant over the entire range of collagen concentrations. 10 γ =γc γ =0.9γc γ =1.1γc rangements are deviations from the imposed affine de- formation and are apparent as divergent strain fluctua- tions as shown in Fig. 6. In the thermodynamic limit of W → ∞, the non-affine rearrangements in the network grow without bound giving rise to the divergent time scale of energy relaxation. 100 10-1 10-2 0 H / H 10-3 1.4 1.2 10-4 α 1.0 10-5 10-6 101 0.8 0.6 2.8 3.0 3.2 3.4 (cid:1)z(cid:0) 3.6 3.8 4.0 103 t/t0 104 105 102 Figure 9: (Color online) Energy versus time for two different network connectivities, hzi ≃ 3.6 (red) and hzi ≃ 3.2 (black) for simple shear. The bending rigidity κ is set to 0 to take only the central-force interactions into account. Energy is ex- pressed in units of the initial energy in the network just after the affine deformation, H0. The time is expressed in arbitrary units, chosen to be the same for both connectivities. For both connectivities, elastic energy stored in the network decays as a function of time. Very close to the critical strain, the re- laxation dynamics follow a power-law E(t) ∼ t−1 indicated with the thick blue line. The exponent α for critical slowing down is insensitive to the network connectivity in the range of 3.0-3.8 as shown in the inset. rounding solvent can be modeled in a simple Stokesian fashion and can be lumped on the network node. This is admittedly a highly simplified version of network dy- namics. We subjected central-force subisostatic networks with connectivities in range of 3.0 − 3.8 to an affine shear of γ = γc, 0.9γc and 1.1γc. The network is floppy for γ ≤ γc implying that the total elastic energy stored in the network decays to zero in the long-time limit. Since the network is rigid for γ > γc, the total elastic energy should relax to a finite value after a characteristic relax- ation time. In Fig. 9, we show the time evolution of the total elastic energy stored in the network for two con- nectivities hzi = 3.2 and 3.6. Clearly, for γ ≶ γc, there is a characteristic relaxation time. However, at γ = γc the slowed down dynamics are robustly captured in the power law scaling of the total elastic energy in the net- work as a function of time. For longer times, the elastic energy stored in the network decays as E(t) ∼ t−α at γ = γc with α ≃ 1 implying that the relaxation time scale is divergent. This inverse-time decay is apparent in all the connectivities considered in this study. Unlike the critical exponents f and φ, the exponent associated with critical slowing down does not evolve with connectivity. The divergent time scale of relaxation at the critical point has its origin in the highly delocalized structural rearrangements in the network. These structural rear- VII. DISCUSSION AND CONCLUSIONS In this study, we focus on the mechanical critical be- havior in fiber networks. The networks considered are athermal, disordered, and are by construction, subiso- static. The criticality is driven by the applied global deformation and is the fundamental mechanism of the nonlinear mechanics of such networks. Unlike the iso- static connectivity threshold which depends on the pre- cise balance of the number of constraints to the degrees of freedom, any generic subisostatic network exhibits criti- cal behavior when subjected to an external deformation. The criticality is evident in the neighborhood of a strain that is determined by the network architecture. One of the hallmark features of critical phenomena is the power-law scaling of the order parameter in the vicinity of the critical point. We show that the stiffness of subisostatic networks with central-force interactions scales as a power-law, K ∼ ∆γf , where ∆γ = γ −γc ≥ 0 is the distance measured from the critical strain and f is a critical exponent. Additional interactions such as resistance to bending stabilize subisostatic networks in the subcritical regime ∆γ < 0 such that for γ << γc, K ∼ κ where κ is the bending rigidity. From the per- spective of a critical phenomenon, finite bending rigidity can be considered as an auxiliary field that suppresses the strain-driven criticality. For κ > 0 the stiffness at the critical strain is finite and depends in a power-law fashion on the strength of bending and stretching inter- actions. Drawing analogy with the ferromagnetic phase transition, where H, the applied magnetic field is the auxiliary field, we capture the crossover of stiffness from bend-dominated to stretch-dominated regimes in terms of a universal scaling function. Another important signature of criticality besides the power-law scaling of the order parameter is the diver- gence of fluctuations in the order parameter at the crit- ical point. In athermal subisostatic networks, the order parameter K is zero at the critical strain and exhibits no fluctuations. However, on considering the deviation of the strain field within the network from the globally imposed affine field, one can create measures for fluc- tuations. We construct one such measure: the strain- derivative of average bending-angle in the network and using finite size scaling demonstrate its divergence in the thermodynamic limit. Recently Xu et. al have developed an image analysis software SOAX which can accurately track fibers in 3D [54]. It is an interesting idea to use SOAX together with confocal shear cell rheology [55] to experimentally measure the average bending angle in re- constituted biopolymer networks. We also study a highly simplified model of network dynamics to test if the network relaxation at the critical point exhibits signatures of critical slowing down. We subject subisostatic networks to an affine shear and study the relaxation of the total elastic energy in the network as a function of time. We find that the elastic energy decays as a power-law in time as ∼ t−1 at the critical strain. The power-law decay implies a divergent relaxation time at the critical strain. We find that the dynamics of networks prepared over a wide range of connectivity hzi = 3.0−3.8 remain the same, i.e., the critical exponent associated with slowing down at the critical strain appears to be insensitive to the connectivity in the network. The analogy with the ferromagnetic phase transition guides us in writing an approximate equation for the scaling function that captures the crossover of stiffness from bend-dominated to stretch-dominated regimes. We demonstrate that the derived equation is highly accu- rate in describing the entire nonlinear stiffness vs. strain curves for any bending rigidity. Since concentration in experiments can be mapped to the reduced bending rigid- ity in our network model, the equation for the crossover function can equivalently describe the stiffness vs. strain curves for any concentration of the protein in the exper- iments. We show that the equation accurately describes the stiffness of collagen networks with a single fit pa- rameter. The excellent agreement of model predictions with the experiments provides strong evidence for crit- icality as the underlying mechanism of the well known 11 nonlinear mechanics of athermal fibrous networks such as collagen [1, 33, 40, 41] and bundled actin [56 -- 58]. A surprising observation is that under simple shear, the critical exponents f and φ appear to be independent of the spatial dimensionality. This is a highly intriguing and also puzzling observation. The critical exponents, as is known from the theory of critical phenomena, depend on the spatial dimensionality. However, the exponents are not constant as they change with the average connectivity in the network. The variation of critical exponents along a critical line is similar to the Ashkin-Teller and 8-vertex models [49 -- 51]. The variation in the exponents occurs over a range of connectivities that is significantly larger than that found in collagen networks. Therefore, one can use a unique set of exponents, f ≃ 0.8 and φ ≃ 2.1 to describe the me- chanics of collagen networks prepared at different concen- trations [1]. The uniqueness of the exponents also allows us to relate the two characteristic strains of a subisostatic network, onset of stiffening strain and critical strain via the critical exponents as γ0 ∼ γφ−f . c In sum, the mechanics of disordered fibrous networks can be understood within the framework of an athermal strain-driven critical phenomenon. The mechanical crit- icality is a generic phenomenon exhibited by all subiso- static networks. We apply our model to collagen net- works which are ubiquitous in biology and find strong ev- idence for the idea that mechanical critical behavior un- derlies the strain-stiffening response of collagenous net- works [1] A. Sharma, A. Licup, K. Jansen, R. Rens, M. Sheinman, G. Koenderink, and F. MacKintosh, Nature Physics 12, 584 (2016). ical Review Letters 101, 215501 (2008). [15] W. G. Ellenbroek, Z. Zeravcic, W. van Saarloos, and M. van Hecke, Europhysics Letters 87, 34004 (2009). [2] P. Fratzl, Collagen: structure and mechanics (Springer [16] C. P. Broedersz, X. Mao, T. C. Lubensky, and F. C. Science & Business Media, 2008). [3] J. C. Maxwell, Philosophical Magazine 27, 294 (1864). [4] A. J. Liu and S. R. Nagel, Nature 396, 21 (1998). [5] T. S. Majmudar, M. Sperl, S. Luding, and R. P. MacKintosh, Nature Physics 7, 983 (2011). [17] M. Sheinman, C. P. Broedersz, and F. C. MacKintosh, Physical Review E 85, 021801 (2012). [18] A. J. Licup, A. Sharma, and F. C. MacKintosh, Phys. Behringer, Physical Review Letters 98, 058001 (2007). Rev. E 93, 012407 (2016). [6] M. Van Hecke, Journal of Physics: Condensed Matter [19] D. A. Head, A. J. Levine, and F. C. MacKintosh, Physical 22, 033101 (2010). [7] W. van Saarloos, M. Wyart, A. J. Liu, and S. R. Nagel, The jamming scenario: An introduction and outlook (2010). Review Letters 91, 108102 (2003). [20] J. Wilhelm and E. Frey, Physical Review Letters 91, 108103 (2003). [21] M. Sheinman, C. Broedersz, and F. MacKintosh, Physi- [8] M. F. Thorpe, Journal of Non-Crystalline Solids 57, 355 cal review letters 109, 238101 (2012). (1983). [22] M. Dennison, M. Sheinman, C. Storm, and F. C. MacK- [9] S. Feng and P. N. Sen, Physical Review Letters 52, 216 intosh, Physical review letters 111, 095503 (2013). (1984). [10] D. J. Jacobs and M. F. Thorpe, Physical Review Letters [23] S. Alexander, Physics Reports 296, 65 (1998). [24] C. P. Broedersz and F. C. MacKintosh, Soft Matter 7, 75, 4051 (1995). 3186 (2011). [11] M. Latva-Kokko, J. Mäkinen, and J. Timonen, Physical [25] C. P. Broedersz, M. Sheinman, and F. C. MacKintosh, Review E 63, 046113 (2001). Physical Review Letters 108, 078102 (2012). [12] P. Olsson and S. Teitel, Physical Review Letters 99, [26] X. Mao, O. Stenull, and T. C. Lubensky, Physical Review 178001 (2007). E 87, 042602 (2013). [13] D. A. Head, Physical Review Letters 102, 138001 (2009). [14] M. Wyart, H. Liang, A. Kabla, and L. Mahadevan, Phys- [27] D. A. Head, A. J. Levine, and F. C. MacKintosh, Physical Review E 68, 061907 (2003). [28] E. Huisman, T. Van Dillen, P. Onck, and E. Van der Giessen, Physical review letters 99, 208103 (2007). [29] A. Shahsavari and R. Picu, Physical Review E 86, 011923 ersz, W. Messner, F. Nakamura, T. P. Stossel, F. C. MacKintosh, and D. A. Weitz, Physical Review E 79, 041928 (2009). 12 (2012). [30] E. Conti and F. C. MacKintosh, Physical Review Letters 102, 088102 (2009). [31] A. Sharma, M. Sheinman, K. M. Heidemann, and F. C. MacKintosh, Physical Review E 88, 052705 (2013). [32] K. M. Heidemann, A. Sharma, F. Rehfeldt, C. F. Schmidt, and M. Wardetzky, Soft matter 11, 343 (2015). [33] A. J. Licup, S. Munster, A. Sharma, M. Sheinman, L. M. Jawerth, B. Fabry, D. A. Weitz, and F. C. MacKintosh, Proceedings of the National Academy of Sciences 112, 9573 (2015). [34] D. A. Head, F. C. MacKintosh, and A. J. Levine, Physical Review E 68, 025101 (2003). [35] A. S. van Oosten, M. Vahabi, A. J. Licup, A. Sharma, P. A. Galie, F. C. MacKintosh, and P. A. Janmey, Scien- tific reports 6, 19270 (2016). [36] M. Vahabi, A. Sharma, A. J. Licup, A. S. van Oosten, P. A. Galie, P. A. Janmey, and F. C. MacKintosh, Soft matter 12, 5050 (2016). [37] M. Achilli and D. Mantovani, Polymers 2, 664 (2010). [38] M. S. Hall, R. Long, X. Feng, Y. Huang, C.-Y. Hui, and M. Wu, Experimental cell research 319, 2396 (2013). [39] R. Picu, Soft Matter 7, 6768 (2011). [40] S. Motte and L. J. Kaufman, Biopolymers 99, 35 (2013). [41] I. K. Piechocka, A. S. van Oosten, R. G. Breuls, and G. H. Koenderink, Biomacromolecules 12, 2797 (2011). [42] J. P. Straley, Journal of Physics C: Solid State Physics 9, 783 (1976). [43] H. Hatami-Marbini and R. Picu, Physical Review E 77, 062103 (2008). [44] A. Arrott and J. E. Noakes, Physical Review Letters 19, 786 (1967). [45] N. Goldenfeld, Lectures on phase transitions and the renormalization group (Addison-Wesley, Advanced Book Program, Reading, 1992). [46] C. P. Broedersz and F. C. MacKintosh, Reviews of Mod- ern Physics 86, 995 (2014). [47] D. Stauffer and A. Aharony, Introduction to percolation theory (CRC press, 1994). [48] R. Rens, M. Vahabi, A. Licup, F. MacKintosh, and A. Sharma, Journal of Physical Chemistry B (2016). [49] J. Ashkin and E. Teller, Physical Review 64, 178 (1943). [50] R. J. Baxter, Physical Review Letters 26, 832 (1971). [51] L. P. Kadanoff and A. C. Brown, Annals of Physics 121, 318 (1979). [52] N. P. Butch and M. B. Maple, Physical review letters 103, 076404 (2009). [53] D. Fuchs, M. Wissinger, J. Schmalian, C.-L. Huang, R. Fromknecht, R. Schneider, and H. v. Löhneysen, Phys- ical Review B 89, 174405 (2014). [54] T. Xu, D. Vavylonis, F.-C. Tsai, G. H. Koenderink, W. Nie, E. Yusuf, I.-J. Lee, J.-Q. Wu, and X. Huang, Scientific reports 5 (2015). [55] R. C. Arevalo, P. Kumar, J. S. Urbach, and D. L. Blair, PLOS one 10, e011802 (2015). [56] M. L. Gardel, J. H. Shin, F. C. MacKintosh, L. Mahade- van, P. A. Matsudaira, and D. A. Weitz, Science 304, 1301 (2004). [57] C. Storm, J. Pastore, F. C. MacKintosh, T. C. Lubensky, and P. A. Janmey, Nature 435, 191 (2005). [58] K. E. Kasza, G. H. Koenderink, Y. C. Lin, C. P. Broed-
1701.01693
1
1701
2017-01-06T17:02:21
Emergent stochastic oscillations and signal detection in tree networks of excitable elements
[ "physics.bio-ph", "cond-mat.dis-nn", "nlin.AO", "q-bio.NC" ]
We study the stochastic dynamics of strongly-coupled excitable elements on a tree network. The peripheral nodes receive independent random inputs which may induce large spiking events propagating through the branches of the tree and leading to global coherent oscillations in the network. This scenario may be relevant to action potential generation in certain sensory neurons, which possess myelinated distal dendritic tree-like arbors with excitable nodes of Ranvier at peripheral and branching nodes and exhibit noisy periodic sequences of action potentials. We focus on the spiking statistics of the central node, which fires in response to a noisy input at peripheral nodes. We show that, in the strong coupling regime, relevant to myelinated dendritic trees, the spike train statistics can be predicted from an isolated excitable element with rescaled parameters according to the network topology. Furthermore, we show that by varying the network topology the spike train statistics of the central node can be tuned to have a certain firing rate and variability, or to allow for an optimal discrimination of inputs applied at the peripheral nodes.
physics.bio-ph
physics
Emergent stochastic oscillations and signal detection in tree networks of excitable elements Justus Kromer1, Ali Khaledi-Nasab2, Lutz Schimansky-Geier3,4, and Alexander B. Neiman2,5,* 1Center for Advancing Electronics Dresden, TU Dresden, Mommsenstrasse 15, 01069 Dresden, Germany 2Department of Physics and Astronomy, Ohio University, Athens, Ohio 45701, USA 3Department of Physics, Humboldt-Universitat zu Berlin, Newtonstrasse 15, 12489 Berlin, Germany 4Bernstein Center for Computational Neuroscience, Berlin, Germany 5Neuroscience Program, Ohio University, Athens, Ohio 45701, USA *[email protected] ABSTRACT We study the stochastic dynamics of strongly-coupled excitable elements on a tree network. The peripheral nodes receive independent random inputs which may induce large spiking events propagating through the branches of the tree and leading to global coherent oscillations in the network. This scenario may be relevant to action potential generation in certain sensory neurons, which possess myelinated distal dendritic tree-like arbors with excitable nodes of Ranvier at peripheral and branch- ing nodes and exhibit noisy periodic sequences of action potentials. We focus on the spiking statistics of the central node, which fires in response to a noisy input at peripheral nodes. We show that, in the strong coupling regime, relevant to myeli- nated dendritic trees, the spike train statistics can be predicted from an isolated excitable element with rescaled parameters according to the network topology. Furthermore, we show that by varying the network topology the spike train statistics of the central node can be tuned to have a certain firing rate and variability, or to allow for an optimal discrimination of inputs applied at the peripheral nodes. Introduction Coupled noisy excitable systems serve as relevant models for a wide range of natural phenomena, including pattern formation in chemical reactions1, 2 and in social networks3 -- 6, dynamics of gene regulatory networks7 and of single and networked neurons8 -- 10. Networks of noisy excitable elements exhibit a rich variety of spatio-temporal dynamics, depending on the strength and topology of coupling and the noise intensity11 -- 13. For example, the coherence of emergent network oscillations can be controlled by modifying the noise intensity, the coupling strength, or by changing the network size or topology14 -- 19. The dynamic range and sensitivity of complex networks of excitable elements to external stimuli can by optimized for critical topologies20 -- 22. In the present paper, we focus on the dynamics of regular tree networks of strongly coupled excitable elements which receive random and independent excitations to their peripheral nodes, as sketched in Fig. 1. Our study is motivated by the morphology of certain peripheral sensory neurons, which possess branched myelinated dendritic terminals at their receptive fields, with multiple nodes of Ranvier. Their extended terminal branching resembles the dendrite structure of neurons in the central nervous system (CNS)23, 24. Myelinated segments form a tree-like structure with nodes of Ranvier at each branching point. Myelination terminates at peripheral nodes of Ranvier, called heminodes, which receive sensory signals. Thus, such sensory neurons may possess multiple spike initiation zones at heminodes which encode a local sensory signal into a stream of action potentials (APs) which are merged into a single output spike train transmitted to the CNS25. Examples for such neurons are the afferent innervation of muscle spindles26 -- 29, pain receptors30, cutaneous mechanoreceptors31, 32, and lung receptors33. Interestingly, sensory neurons with myelinated dendrites may exhibit spontaneous activity characterized by coherent periodic spiking, despite that their peripheral heminodes presumably receive uncorrelated noisy excitations28. Figure 1 then can be viewed as a model for a branched myelinated dendritic terminal, where peripheral nodes receive uncorrelated stochastic inputs and are linked by myelinated segments. Due to the high density of Na+ ion channels at the nodes of Ranvier, APs may be excited independently at different peripheral nodes. High electrical conductivity of myelinated segments, which link the individual nodes, result in a strong coupling between the nodes. Therefore, their stochastic dynamics synchronizes. This may result in noisy periodic spiking of the primary branching (central) node, as we have shown for star networks of excitable elements34. Here we use a biophysical model for nodes of Ranvier connected by myelinated links on regular trees and show numerically 3 2 1 0 Figure 1. Tree network with branching order, d = 3, and G = 3 generations. Peripheral nodes are marked red and receive external excitations. Dashed circles indicate corresponding shells of the tree's generations, g = 1, 2 and 3; generation g = 0 refers to the central (primary branching) node (green). For a discrete cable model of myelinated dendrites, active elements are nodes of Ranvier, which are connected by passive resistors. and analytically that the collective response of the network can be deduced from the stochastic dynamics of a single effective node with parameters scaled according to the network size and topology. Thus, our study allows for the prediction of the stochastic network dynamics from the tree topology. We then discuss how the tree topology affects the firing statistics of the central node and the discriminability of input signals. Model and Methods Discrete Cable Model In the present paper, we study the stochastic dynamics of excitable elements linked on a regular tree (see Fig. 1). Branching starts at the primary (central) node (number 0 in Fig. 1) and continues through several generations. Only the peripheral nodes receive external inputs. Referring to a model of branched myelinated dendrites, these peripheral nodes are called heminodes and receive inputs from thin unmyelinated processes (neurites). APs are initiated at the heminodes and then propagate on the tree towards the primary branching node and eventually to the CNS. Here we consider regular trees whose topology is characterized by two parameters: the branching, d, and the number of generations, G. Given these two parameters, the total number of nodes, N, and the number of peripheral nodes, Np, are given by N = dG+1 − 1 d − 1 and Np = dG, (1) respectively. The dynamics of the membrane potential is approximated by a discrete cable model35 in which nodes of Ranvier are connected by passive resistive links according to the network topology. All active nodes and passive links are assumed to be identical, except that peripheral nodes receive external inputs. The membrane potential Vk(t) of the kth node obeys the dynamics C Vk = −Iion[Vk,uk] + k uk = a (Vk)(1− uk)− b (Vk)(1− uk), j=0 N−1 Ak, j(Vj −Vk) + Iextk , (2) where the index k = 0,1,2, ...,N − 1 marks the respective node. In particular, k = 0 refers to the central node. In Eq. (2) the term Iion[Vk,uk] stands for nodal ionic currents and uk(t) is a vector whose components are the gating variables of the nodal 2/15 (cid:229) ion channels and C = 2 m F/cm2 is the nodal capacitance per area. In the following we use two particular models for the nodes of Ranvier: a Hodgkin-Huxley-type (HH) model with Na+ and leak currents34 and the Frankenhaeuser-Huxley (FH) model which includes additional K+ and persistent Na+ currents. The HH nodal model includes two gating variables, m and h, for Na+ channels, i.e. Iion[V,u] = Iion[V,m,h]. The FH model includes two additional gating variables, n for K+, and p for persistent Na+ channels: Iion[V,u] = Iion[V,m,h,n, p]. The detailed equations and parameters of the nodal models are provided in the Supplementary Material. The coupling term in Eq. (2), k (cid:229) N−1 j=0 Ak, j(Vj − Vk), contains the adjacency matrix A of the tree graph and the coupling strength, k , in units of Siemens per area. Its value can be calculated from the sizes of the node and myelinated links, and the axoplasmic resistivity: k = a 4lLr , (3) where a is the diameter of the node (and of links), l is the nodal length, L is the length of connecting links and r is the axoplasmic resistivity. For example, for r = 100 W cm, the nodal diameter and length a = 10 m m, l = 1 m m, and the length of myelinated segment L = 200 m m, the coupling strength is k = 1250 mS/cm2. This provides a biophysically-plausible range for k , which we use as a control parameter in the following. The external current Iext is applied only to the peripheral nodes and consists of a constant part I and noisy part, i.e. Iextk = d k,p[I +√2Dx p(t)], (4) where p denotes indicies of peripheral nodes; d k,p is the Kronecker delta; D scales the intensity of the Gaussian white noise x p(t), which is uncorrelated for different peripheral nodes, hx i(t)x j(t +t )i = d i, j d (t ). Thus, peripheral nodes receive random uncorrelated inputs. Equations (2) were integrated numerically using explicit Euler -- Maruyama methods with timestep of 0.1m s. Variability of Generated Sequences of Action Potentials Our primary interest is the statistics of a spike train generated by the central node. A spike is identified as a full-size AP with a magnitude of at least 60 mV. We extracted a sequence of spike times, t j, at the central node from 60 -- 120 s long simulation runs. The corresponding sequence of interspike intervals (ISIs) D t j = t j+1 − t j, is characterized by the mean firing rate, r and the coefficient of variation, CV as, r = 1 t ji hD , CV = rqh(D t j −hD t ji)2i, where the average is taken over all ISIs in the spike train of the central node. (5) Signal Detection To characterize the signal detection capacity of a tree network, we considered a small constant stimulus, D I, applied to the peripheral nodes in addition to the stimulus I, and calculated a normalized distance between resulting spike count distributions of the central node with and without this addition. Such a measure of distance is given by the discriminability, d′, defined as36, d′ = 2m T (I)− m T (I + D I) s T (I) + s T (I + D I) , (6) where m T and s T are the mean and standard deviation of the spike count in a time interval T , respectively. The discrim- inability quantifies how well the network responses to two different stimuli, I and I + D I, can be distinguished by observing corresponding spike count statistics at the central node. The discriminability is related to the Fisher information, which provides the theoretical limit of how accurately a stimulus I can be estimated by observing a spike train37. For the spike count statistics, a lower bound of the Fisher information can be written as36, JLB(I) = 1 s 2 T (I) (cid:18) dm T dI (cid:19)2 , and is related to the discriminability, d′ by36, d′ ≈ D IpJLB(I). (7) (8) 3/15 Larger values of the Fisher information refer to more accurate estimation of the stimulus from the spike train and so to better discrimination between two stimuli I and I + D I. The discriminability Eq. (6) was calculated by collecting spike counts of the central node for 5000 independent time intervals of lengths T = 200 ms, and calculating the mean and standard deviation for two values of the stimulus, I and I + D I, applied to the peripheral nodes of a tree network36. We also calculated the lower bound of the Fisher information Eq. (7) for the single uncoupled node as a function of the input current (stimulus) I and the noise intensity, D, using a similar numerical procedure. Results Emergence of Periodic Firing in Deterministic Tree Networks At first, we consider the case of a deterministic input, D = 0. In the absence of the external input, Iext = 0, an isolated node is in the excitable regime. A sufficiently high constant current, IAH results in a subcritical Andronov-Hopf bifurcation of the equilibrium state rendering an isolated node to fire a periodic sequence of APs. The corresponding limit cycle disappears in a saddle-node bifurcation for a lower external current, ISN. For the HH nodal model the saddle-node bifurcation occurs at ISN ≈ 28.15m A/cm2 and the subcritical Andronov-Hopf bifurcation at IAH ≈ 29.06m A/cm2, so in a narrow range ISN < I < IAH an isolated node is bistable, possessing a stable equilibrium and a stable limit cycle. When the nodes are coupled on a tree network and external currents are applied to the peripheral nodes, the dynamics of the network may become quite complex. For example, in case of weak coupling, peripheral nodes fire APs, which fail to propagate to the central node, so that nodes in the inner generations of the network exhibit small-amplitude spikes. For a stronger coupling, nodes in the inner generations may fire APs, but with skipping relative to APs in the periphery, demonstrating various m : n synchronization patterns. However, for strong coupling and sufficiently high external currents the network shows fully synchronized periodic firing. A comprehensive analysis of the deterministic dynamics is beyond the scope of this study. Instead, since our primary interest is in the emergence of periodic sequences of full-size APs at the central node, we address the following question: Given the tree topology, G and d, and the coupling strength, k , what is a threshold value Ith of a constant current applied to peripherals, I, which makes the central node to generate repetitive firing of full-size APs? To this end, we perform simulations of tree networks with given k , G and d. Initially membrane potentials of the individual nodes are randomly distributed around the stable equilibrium of an isolated node for I = 0. Then we apply a current I > 0 and determined the minimal value, Ith of I at which the central node generated APs repetitively at steady state. Results are shown in Fig. 2. At the central node N S I / I h t (a) d = 2 2.0 1.8 1.6 1.4 1.2 1.0 N S I / I h t (b) d = 3 1.5 1.4 1.3 1.2 1.1 1.0 G = 1 G = 2 G = 3 G = 4 0.1 1 10 N(mS/cm2) 100 1000 0.1 1 10 N(mS/cm2) 100 1000 Figure 2. Threshold constant current Ith vs the coupling strength, k , which gives rise to repetitive APs at the central node of the tree network with branching d and the number of generation, G. The threshold current is normalized by the bifurcation value ISN = 28.15 m A/cm2 at which the single isolated HH node starts to generate a periodic sequence of APs. Colors (see legends) refer to the indicated number of generations, G. Panels correspond to trees with branching d = 2 (a) and d = 3 (b), respectively. periodic firing of APs occurs for values of I and k above the corresponding curves in Fig. 2. Below these curves, the network is excitable in the sense that no repetitive firing of APs is observed at the central node. In the following, we refer to these two regimes as oscillatory (repetitive firing of full-size APs by the central node) and excitable (no repetitive firing of APs by the central node). The threshold value of the external current, Ith, increases for weak and moderate values of the coupling strength. Consequently the network needs stronger external input to the peripheral nodes to sustain periodic firing of the central node. Figure 2 shows two distinct coupling regimes. For weak coupling, k < 2 mS/cm2, the threshold current Ith is independent of the network size, i.e. the number of generations, G, and branching, d. In contrast, for strong coupling, k > 60 mS/cm2, the threshold current saturates, and its value increases with increasing number of generations. This is illustrated further in Fig. 3 showing the threshold current vs the number of generations for strong coupling. Note that the strong coupling regime spans 4/15 the range of realistic coupling strengths for models of branched myelinated dendrites. As can be seen in Fig. 3, the threshold current follows a characteristic dependence saturating for trees with a large number of generations, G, and decreases with the increase of branching, d. Apparently, this dependence follows the scaling relation: Ith/Ib = R−1(G,d), (9) where Ib is a bifurcation value of the constant current in the isolated single node and the scaling factor R(G,d) is the ratio of the number of peripheral nodes to the total number of nodes, R(G,d) = Np N = dG(d − 1) dG+1 − 1 , (10) whith Np = dG and N = (cid:229) G derived below. Deterministic trees with the FH nodal model show similar dynamics with the same scaling as in Fig. 3. k=0 dk, for regular trees. This scaling relation, Eq. (9), holds for strongly-coupled trees and is N S h t I / I 2.00 1.75 1.50 1.25 d = 2 d = 3 1 2 3 4 5 6 7 8 G Figure 3. Threshold constant current Ith vs. the number of generations for strong coupling k = 100 mS/cm2 and indicated values of branching, d. The threshold current is normalized by the bifurcation value ISN = 28.15 m A/cm2 at which a single isolated HH node starts to fire a periodic sequence of APs. Symbols indicate results from simulations and solid lines show the scaling relation Eq. (9). Stochastic Dynamics The addition of uncorrelated noise to the peripheral nodes allows for the generation of APs in the excitable regime. Fig. 4 shows an example of the stochastic dynamics for a tree with G = 5 generations and d = 3 branching. In the excitable regime (I = 20 m A/cm2) noise of sufficient intensity induces APs in peripheral nodes. For weak coupling (k = 0.3 mS/cm2) noise-induced APs in adjacent generations are not synchronized (superimposed spikes for peripheral nodes fill densely corre- sponding generation panels) and do not propagate beyond the 2-nd generation, which shows only sparse APs. Increasing the coupling strength leads to progressive synchronization of nodes in adjacent generations and finally results in the generation of APs in the central node. For strong coupling the whole network fires almost in synchrony. We note, however, that even for strong coupling outer generations show some spike jitter. We also note that strong coupling leads to slower and more random firing of APs. As observed for star networks34, the dynamics of the central node in a tree network depends non-monotonously on the coupling strength. As shown in Fig. 5, there exist optimal, rather small values of the coupling strength for excitable and oscillatory trees at which fastest (maximum firing rate) and most coherent (minimal coefficient of variation, CV ) firing is observed, respectively. For extremely weak coupling APs, which are fired by different peripheral nodes, are not synchronized and fail to propagate to the central node (Fig. 4, upper left panel). Increasing the coupling strength leads to stronger interaction between the branch nodes and results in synchronous firing of all nodes. However, the size of a tree, i.e. the number of generations, is critical for the firing statistics of the central node. Further- more, excitable and oscillatory trees demonstrate qualitatively different behaviour in the biologically-relevant strong coupling regime. In excitable trees, firing of APs becomes slower and more irregular if the coupling is strengthened and trees with more generations are considered. For large G and strong coupling firing stops [Fig. 5(a1)] since excitatory inputs to peripheral nodes are too weak to sustain firing of APs. In contrast, in oscillatory trees, the firing rate saturates for strong coupling [Fig. 5(a2)] and firing becomes more regular if strongly-coupled trees with more generations are considered [Fig. 5(b2)]. 5/15 κ = 0.3 κ = 1.0 κ = 10 5 4 3 2 1 0 5 4 3 2 1 0 20 ms κ = 80 100 mV Figure 4. Voltage traces of the HH nodes for a tree with G = 5 generations and d = 3 branching in the excitable regime. Peripheral nodes were excited by the external currents (4) with parameters I = 20 m A/cm2 and D = 500 (m A/cm2)2ms. Each panel shows 200 ms long superimposed voltage traces of nodes within a generation g, g = 1, ..,5, for the indicated coupling strength, k (in mS/cm2). Horizontal axis is time. Numbers next to voltage traces indicate generations within the tree, g = 0 corresponds to the central node and g = 5 corresponds to the peripheral generation, respectively. Scaling of Effective Current and Noise intensity In the strong coupling regime the dynamics of the central node of a tree network can be described by the dynamics of a single isolated node with membrane potential V0(t) and with effective input current Ieff and an effective Gaussian noise with intensity Deff, i.e. the influence of the coupling term on the dynamics can be approximated by a constant current and a white Gaussian noise. Then the dynamics of the membrane potential of the central node in eq. (2) can be approximated by C V0 ≈ f [V0,u0] + Ieff +p2Deff x eff(t), u0 ≈ g[V0,u0]. In the following, we derive those effective parameters for regular trees of diffusively-coupled nodes. In the network model, the dynamics of the membrane potential of k-th node is given by C Vk = f [Vk,uk] + k N−1 j=0 Ak, j(Vj −Vk) + Iextk , k = 0,1, ...,N − 1. (11) (12) In order to derive approximations for the scaling of the effective current Iext and the noise intensity Deff, we extend the approach of Kouvaris et al.38, who considered the propagation of excitable waves in a tree network of identical Fitz-Hugh Nagumo nodes in the absence of noisy inputs. Following their approach, we consider the dynamics of the average membrane potential hVig (termed density by Kouvaris et al.) in each shell in a tree. Here and in the following hig denotes averaging over all nodes of the gth shell, hVig := 1 dg Vk. gth shell (13) The dynamics of those densities can be obtained by averaging the respective equations for the dynamics of the membrane potentials, Eq. (12), over all nodes in one shell. Since the total number of connections between nodes in shell g and g− 1 is dg, we obtain k d(hVi1 −V0), Ch Vig = h f (V,u)ig + k (hVig−1 − (d + 1)hVig + dhVig+1) , k (hViG−1 −hViG) + I +q2 D  Np g = 0, 0 < g < G, x G(t), g = G. (14) Note that, since peripheral nodes are subject to independent white Gaussian noises, the corresponding equation for the aver- aged membrane potentials of the peripheral generation contains white Gaussian noise x G(t) with reduced intensity D Np . 6/15 (cid:229) (cid:229) G = 1 G = 3 G = 5 100 (a2) ) z H ( > f < 80 60 40 20 0 100 (a1) ) z H ( > f < 80 60 40 20 0 1.0 0.8 V C 0.6 0.4 0.2 0.0 0.1 1 (b1) 10 N(mS/cm2) 100 1000 0.1 1 10 N(mS/cm2) 100 1000 (b2) 1.0 0.8 V C 0.6 0.4 0.2 0.0 G = 1 G = 3 G = 5 0.1 1 10 N(mS/cm2) 100 1000 0.1 1 10 100 1000 N(mS/cm2) Figure 5. Mean firing rate (a1, a2) and coefficient of variation of interspike intervals (b1, b2) of the central node versus coupling strength for tree networks with d = 2 branching for the indicated numbers of generations, G. Left panels (a1, b1) correspond to the excitable regime with the constant current I = 20 m A/cm2; right panels (a2, b2) refer to the oscillatory regime with I = 60 m A/cm2. The noise intensity is D = 500 (m A/cm2)2ms. Since the coupling terms depend only on the difference between densities of the membrane potentials in adjacent genera- tions D Vg := hVig −hVig+1, we consider the dynamics of those differences next. Subtracting equations for h Vig yields, CD Vg = (h f (V,u)ig −h f (V,u)ig+1) + D Ig + D x g(t) + g = 0, k (dD V1 − (d + 1)D V0) , k (D Vg−1 − (d + 1)D Vg + dD Vg+1) , 0 < g < G− 1, k (D VG−2 − (d + 1)D VG−1) , g = G− 1,   (15) where D Ig = −d g,G−1 I and D Next, we consider the case of strong coupling. In that case, D Vg becomes small, and the membrane potentials of individual nodes approach the average potentials of the corresponding shell. Thus, we can approximate h f (V,u)ig −h f (V,u)ig+1 by a g D Vg + h.o.. It then follows for strong coupling, Taylor expansion around D Vg = 0, i.e. h f (V,u)ig −h f (V,u)ig+1 ≈ D f 1 i.e. x g(t) = −d g,G−1q2 D x G(t) is Gaussian white noise. g + D f 0 Np k ≫ D f 0 g = 0, f 1 g , g = 0,1, ...,G− 1, (16) that the dynamics of the averaged potential is dominated by the coupling term and D Vg can be approximated by a multidimen- sional Ornstein−Uhlenbeck process, x (t). (17) C d dt D V ≈ k BD V + D I + D Here we introduced the G-dimensional vectors, D V = (D V0,D V1, ...,D VG−1)T , D I = (D I,D I1, ...,D IG−1)T , x (t) = (D x 0(t),D x 1(t), ...,D x G−1(t))T , and the G× G tridiagonal Toeplitz matrix, 0 d d − (d + 1) B = 1 ... .. − (d + 1) ... 0 − (d + 1)   1 0 ... 0 0 ... 0 d ... ... ... ... 1 − (d + 1)   . (18) 7/15 D D 40 (a1) ) z H ( > f < 30 20 10 0 1 2 1.0 (b1) V C 0.8 0.6 0.4 1 2 3 G 3 G d = 2 d = 3 theory theory 3 G d = 2 d = 3 4 5 theory theory d =2 theory 80 (a2) ) z H ( > f < 70 60 50 40 4 5 1 2 d =2 theory 0.20 (b2) 0.15 V C 0.10 0.05 0.00 4 5 1 2 3 G 4 5 Figure 6. Mean firing rate, h fi, (a1, a2) and coefficient of variation, CV , of interspike intervals (b1, b2) of the central node versus the number of generations of tree networks in the strong coupling regime, k = 1000 mS/cm2 and noise intensity D = 500 (m A/cm2)2ms. Left panels (a1, b1) correspond to an excitable tree with constant current I = 20 m A/cm2; right panels (a2, b2) refer to an oscillatory tree with I = 60 m A/cm2. Symbols (cid:3) and (cid:13) mark results of numerical simulations of the corresponding network with the indicated branching, d; solid lines and symbols × show theoretical scaling predictions. In the strong coupling limit (16), deviations of D V from its mean value decay extremely fast and we can use an adiabatic elimination39 to approximate D V by its mean value plus a white Gaussian noise. Both, the mean voltage difference and the intensity of the Gaussian white noise in the strong coupling limit can be obtained by setting the left-hand side of Eq. (17) to zero. This yields D V ≈ − 1 k B−1 (D I + D x (t)) , (19) where B−1 is the inverse of the matrix B. In order to obtain an approximation for the dynamics of the central node, we can use Eq. (19) to replace V0 −hVi1 by D V0 in Eq. (14) for the central node, g = 0. This yields Ch Vi0 = C V0 = f [V0,u0] + dB−1 (D I + D x (t))1 . (20) Here and in the following the index "1" denotes the first component of a G-dimensional vector. Next, the effective parameters Ieff and Deff can be obtained by comparing Eqs. (20) and (11). This yields the effective input current and the intensity of the effective white Gaussian noise, Ieff = d(cid:0)B−1D I(cid:1)1 , Deff = d(cid:0)B−1D For the special case, considered in this study, that only peripheral nodes are subject to noisy inputs, i.e. D I = (0,0, ...,−I)T , x (t) = (0,0, ...,−p2D/Np x G(t))T , the calculation of the effective parameters Ieff and Deff requires only a single compo- and D nent, (1,G), of the inverse matrix, B−1. Since B is a tridiagonal Toeplitz matrix, we can apply the results of Ref.4 to calculate this component (see Supplemental Material for details on calculations) and find for the effective current, x (t)(cid:1)1 . (21) Ieff = Np N I = R(G,d)I, and for the effective noise intensity, Deff = Np N2 D = R(G,d) N D, (22) (23) where the scaling factor R(G,d) is given by Eq. (10). Investigating the scaling of the effective parameters in more detail, we first note that our theory yields the scaling relation, Eq. (9), observed for the deterministic threshold current in Fig. 3. In fact, the same scaling relation applies to the bifurcation 8/15 Mean firing rate 100 d = 2 3 4 5 10 CV Coef. of variation ( ) d = 2 3 4 5 10 ] s m 2 ) 2 m c / A P ( [ D 10 1 0.1 0.01 10 20 G 40 30 I [PA/cm2] <f > 0 10 20 30 40 50 60 70 80 50 60 10 20 CV 0.001 0.01 0.1 >1 G 40 30 I [PA/cm2] 50 60 Figure 7. Heat maps of the mean firing rate, h fi and coefficient of variation, CV , vs input current, I, and noise intensity, D, for the single isolated HH node obtained from numerical simulations. Circles and magenta lines show the scaling of current, Eq. (22), and noise intensity, Eq. (23), for tree networks with indicated values of branching d and with increasing number of generations, G from 1 (top) to 10 (bottom). For tree networks the input current to the peripheral nodes is I = 60 m A/cm2 and the noise intensity is D = 500 (m A/cm2)2ms. values of I in the deterministic model, e.g. the subcritical Andronov-Hopf bifurcation of the equilibrium or the saddle-node bifurcation of the limit cycles. Second, in Fig. 6, we demonstrate the validity of the theoretical scaling predictions by com- paring results for the mean firing rate and the CV from direct simulation of tree networks with those from a single node (11) with input current and noise intensity scaled according to Eqs. (22) and (23), respectively. As illustrated in Fig. 7, we find an excellent correspondence of both results. This indicates that in the strong coupling limit the response of the network can be predicted from the stochastic dynamics of the effective central node. The statistics of interspike intervals for a single isolated node versus input current parameters, i.e. constant component, I, and noise intensity, D, can be easily computed numerically yielding two-dimensional maps, such as shown in Fig. 7. Then for a given size (number of generations, G) and branching, d, of a tree, the scaled parameters, Eqs. (22) and (23), set an operation point for the tree on the parametric map of a single element. Thus, predictions of the firing statistics of the central node of a tree of strongly-coupled excitable elements can be deduced by superimposing parametric dependencies of Ieff and Deff on the parameters of the network. Figure 7 demonstrates this for trees of strongly-coupled HH nodes in the oscillating regime. In trees with more generations G the operation point is shifted towards smaller currents and lower noise intensities, resulting, for oscillatory trees, in slower and more coherent firing of the tree's central node. Finally, in the strong coupling limit the scaling relations (22, 23) are independent of the particular choice of the nodal model, e.g. they are expected to work for either Hodgkin-Huxley or Frankenhaeuser-Huxley nodal models. Signal Detection The signal detection efficiency of a neuron can be quantified using the discriminability and the Fisher information36, 41 -- 44. In case of our model of coupled excitable elements on a tree, we use these measures to characterize how the tree topology affects its ability to distinguish between two stimuli, I and I + D I, applied to the peripheral nodes. The preceding section showed that in the strong coupling limit, the stochastic dynamics of the network could be predicted from the dynamics of a single node with appropriately scaled parameters of the input current. Thus, we first analyze the lower bound of the Fisher information of a single node. Equation (7) indicates that the Fisher information is determined by two dI (cid:17)2 factors: the term(cid:16) dm T , which is related to the slope of the so-called f − I curve (mean firing rate vs input current curve) and determines the sensitivity of a neuron to small variations of the input current. The sensitivity is largest in the vicinity of the bifurcation point, where the limit cycle is born, and where the slope of the f − I curve is the steepest. In this region, the Fisher information is high. However, the second factor in Eq. (7), the variance of the spike count, may degrade the Fisher information. In the excitable regime, when the input current is below its bifurcation value and APs are induced by noise, the phenomenon of stochastic resonance is observed45, i.e. due to the competition of two factors, the sensitivity and the spike count variance, the Fisher information possesses a maximum at an optimal noise intensity36. Figure 8(a1) shows the lower bound of the Fisher information, JLB, for a single HH node as a function of input current and noise intensity. The Fisher information is maximal for an input current I, which brings the system close to the transition to periodic spiking, i.e. 28 -- 29 m A/cm2. In Fig. 8(a1), a vertical section across the map corresponds to the dependence of the Fisher information on noise intensity. As can be seen, such a dependence is non-monotonous in the excitable regimes, e.g. for I = 15 or I = 20 m A/cm2, indicating the phenomenon of stochastic resonance45, reported before for the original Hodgkin- Huxley neuron model in Ref.36. Indeed, stochastic resonance is a generic phenomenon in excitable systems12, 45 and so the 9/15 I = 20 30 40 55 60 (b1) I = 20 I = 30 I = 40 I = 55 I = 60 1 2 3 G 4 5 6 I = 70 I = 100 I = 115 I = 130 (a1) 100 ] s m 2 ) 2 m c / A m ( [ D 10 1 10 15 (a2) 100 I = ] s m 2 ) 2 m c / A m ( [ D 10 1 30 40 30 25 20 I [PA/cm2] 100 70 35 40 115 130 J 1/2 LB 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 J 1/2 LB 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 y t i l i b a n i m i r c s i d 1.6 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 (b2) y t i l i b a n i m i r c s i d 1.6 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 60 50 I [PA/cm2] 70 80 90 1 2 3 4 G 5 6 Figure 8. Signal detection by a single node and by tree networks. (a): Square root of the lower bound of the Fisher information (7) of the single HH (a1) and single FH (a2) node versus input current and noise intensity. Superimposed white lines with circle symbols show the scaling of the effective input current, Eq. (22), and effective noise intensity, Eq. (23), for tree networks with branching ratio d = 2 and increasing number of generations. Each of these lines corresponds to values of the external current to peripheral nodes, I, indicated at the top of panels (a). The number of generations, G, increases from 1 (top) to 6 (bottom). Noise intensity for tree networks is D = 500 [(m A/cm2)2ms]. (b): Discriminability (6) versus the number of generations, G, for tree networks with branching ratio d = 2 and indicated values of the input current I for HH (b1) and FH (b2) nodes. The increment of the input current is D I = 2 m A/cm2 and noise intensity is D = 500 [(m A/cm2)2ms]. Filled circles refer to results of numerical simulations of corresponding tree networks; solid lines and symbols × show theoretical scaling predictions obtained from simulations of the single HH or FH node with the input current and noise intensity scaled according to Eq. (22) and Eq. (23), respectively. In all panels the spike count statistics was calculated for T = 200 ms windows. The coupling strength on the panels (b1, b2) is k = 1000 mS/cm2. Frankenhaeuser-Huxley (FH) nodal model demonstrates qualitatively similar parameter dependence, shown in Fig. 8(a2). In the absence of noise, a stable equilibrium of the single FH node passes through a subcritical Andronov-Hopf bifurcation at IAH ≈ 60.19 m A/cm2. Consequently, the Fisher information in Fig. 8(a2) is maximal around this value, similar to the HH node. In the excitable regime, e.g. I = 40 m A/cm2, the Fisher information vs. noise intensity passes through a maximum, demonstrating stochastic resonance, again, qualitatively similar to the HH node. The scaling relations for the input current, Eq. (22), and noise intensity, Eq. (23), enable us to predict the signal detection capability of a tree network in the regime of strong coupling. Given the branching, d and the input current to the peripheral nodes, an increase in the number of generations (i.e. tree size) results in a decrease of the effective input current and noise. Then, depending on the particular values of I and D, signal detection by the tree may show distinct dependencies on the tree size, G. This is illustrated in Fig. 8(a1,a2) by superimposing the scaling of the input current and noise intensity on the Fisher information map of the single node. In particular, our theory predicts that in the excitable regime, i.e. when the network does not produce sustained periodic firing in the absence of stochastic inputs, the scaling of I and D may bring an effective operating point of the network across the local maximum of the Fisher information. As can be seen, for instance, for the input current I = 30 or 40 m A/cm2 for HH nodal model, an increase of the number of generations to G =2 -- 4 brings the effective operating point to regions of higher values of the Fisher information; further growth of the tree size eventually suppresses AP firing and thus small signals cannot be detected. In contrast, in the oscillatory regime (e.g. for I = 60 m A/cm2 for HH nodal 10/15 model), the increase of the network size moves the operation point always to regions of higher values of the Fisher information and so the discriminability increases monotonously with the tree size, G. Interestingly, one could predict the input current to the network which for a tree with large enough generations would result in an effective operating point close to bifurcation value of the single node. For example, for the HH nodes, such a value of the external current is I = 55 m A/cm2 and for the FH node, I = 115 m A/cm2. For such currents increasing the tree size should result in a higher degree of signal discrimination. To test these predictions we computed the discriminability (6) for trees with different numbers of generations and for the single node with the scaled values of constant input current and noise intensity according to Eqs. (22, 23). Figure 8(b1,b2) shows excellent correspondence between the respective discriminabilities. Conclusion We have studied the emergence of noisy periodic spiking in regular tree networks of coupled excitable elements. Using bio- physical models of excitable nodes, we showed that noisy periodic network spiking can be generated, although the periphery of the tree is excited by random and independent inputs (Fig. 4). The firing rate and coherence of spiking can be maximized by varying the coupling strength and is altered by changing the network topology (Figs. 5,6). We put special emphasis on the strong coupling regime, which refers to the case of excitable nodes of Ranvier linked by myelinated (dendrite or axon) fibers of a neuron. It is intuitively clear that in the strong coupling limit, the collective dynamics of the network could be described by a single effective excitable system. We have derived the corresponding scaling relations for random inputs Eqs. (22, 23) which allows for reliable predictions of the collective network response based on the stochastic dynamics of a single isolated node with scaled input parameters. Stochastic excitable systems demonstrate non-trivial behaviour versus the noise intensity. Examples include the phenomena of coherence resonance46, whereby the variability of spiking events (e.g. coefficient of variation) is minimal for non-zero noise intensity, and stochastic resonance, characterized by non-monotonous dependence of a response to an external signal on the noise intensity45. Similar phenomena have been observed in networks of excitable elements. In particular, the phenomena of system size stochastic47 and coherence resonance14, which are also observed in strongly-coupled star networks of excitable elements34. As we have shown in the present paper, the phenomenon of system size stochastic resonance also occurs in strongly coupled tree networks, i.e. the number of generations in a tree network of excitable elements can be tuned in order to optimize the network ability to discriminate between different input signals. In particular, our analytical approach allows for the prediction of optimal tree sizes and branching ratio. The analytical approach developed here can be extended to random trees48 in which the branching ratio varies among different generations, yielding similar scaling relations in the strong coupling limit. While we considered networks of identical nodes, our approach can be readily extended to the inhomogeneous case, as long as the condition for strong coupling Eq. (16) is satisfied. Our results suggest a mechanism for the emergence of noisy periodic firing and information coding by peripheral sensory neurons which possess branched tree-like myelinated dendrites28. Such neurons may possess multiple spike initiation zones at peripheral nodes (heminodes) and nodes of Ranvier at branching points. Examples of the muscle spindles27 and cutaneous mechanoreceptors31 indicate that myelinated dendritic trees extend to up to 7 generations. Myelin provides low-resistance links between nodes and fast saltatory conduction of APs, which corresponds to strong coupling between the nodes of Ranvier. For example, the average diameter of a cat muscle spindles afferents ranges from 3 to 13 m m, while links between nodes are relatively short, 50 -- 200 m m27. An estimate of the coupling strength from Eq. (3) yields values well within the range of the strong coupling regime used in our study. The collective noisy periodic firing then may occur due to the synchronized noise-induced generation of APs by stimulating the peripheral heminodes, as described by our model. Given the biophysical properties of the nodes of Ranvier and the sensory inputs, the variability of interspike intervals and the stimulus discrimination capability of a neuron are determined by the ratio of the number of signal-receiving peripheral heminodes to the total number of nodes in the network. Acknowledgements We thank D.F. Russell, E. Schoell, T. Isele for fruitful discussions. AN acknowledges support by the Lobachevsky University of Nizhny Novgorod through the Russian Science Foundation grant 14-41-000440. LSG thanks Ohio University for hospitality and support. Author contributions. AN formulated the problem. JK performed analytical calculation. AKN and AN performed numerical simulations. JK, AKN, LSG and AN wrote and reviewed the manuscript. Competing financial interests. The authors declare no competing financial interests. 11/15 References 1. Kiss, I. Z. & Hudson, J. L. Chemical complexity: Spontaneous and engineered structures. AIChE journal 49, 2234 -- 2241 (2003). 2. Mikhailov, A. S. & Ertl, G. Engineering of Chemical Complexity (World Scientific, 2012). 3. Farkas, I., Helbing, D. & Vicsek, T. Social behaviour: Mexican waves in an excitable medium. Nature 419, 131 -- 132 (2002). 4. Newman, M. E. Spread of epidemic disease on networks. Physical Review E 66, 016128 (2002). 5. Perc, M. Coherence resonance in a spatial prisoner's dilemma game. New Journal of Physics 8, 22 (2006). 6. Borgatti, S. P., Mehra, A., Brass, D. J. & Labianca, G. Network analysis in the social sciences. Science 323, 892 -- 895 (2009). 7. Chen, Y., Kim, J. K., Hirning, A. J., Josi´c, K. & Bennett, M. R. Emergent genetic oscillations in a synthetic microbial consortium. Science 349, 986 -- 989 (2015). 8. Copelli, M., Oliveira, R. F., Roque, A. C. & Kinouchi, O. Signal compression in the sensory periphery. Neurocomputing 65 -- 66, 691 -- 696 (2005). 9. Furtado, L. S. & Copelli, M. Response of electrically coupled spiking neurons: a cellular automaton approach. Physical Review E 73, 011907 (2006). 10. Moldakarimov, S., Bazhenov, M. & Sejnowski, T. J. Feedback stabilizes propagation of synchronous spiking in cortical neural networks. Proceedings of the National Academy of Sciences 112, 2545 -- 2550 (2015). 11. Jung, P. & Mayer-Kress, G. Spatiotemporal stochastic resonance in excitable media. Physical Review Letters 74, 2130 (1995). 12. Lindner, B., Garcıa-Ojalvo, J., Neiman, A. & Schimansky-Geier, L. Effects of noise in excitable systems. Physics Reports 392, 321 -- 424 (2004). 13. Sagu´es, F., Sancho, J. M. & Garc´ıa-Ojalvo, J. Spatiotemporal order out of noise. Reviews of Modern Physics 79, 829 (2007). 14. Toral, R., Mirasso, C. R. & Gunton, J. D. System size coherence resonance in coupled fitzhugh-nagumo models. EPL (Europhysics Letters) 61, 162 (2003). 15. Perc, M. Spatial coherence resonance in excitable media. Physical Review E 72, 016207 (2005). 16. Perc, M. Stochastic resonance on excitable small-world networks via a pacemaker. Physical Review E 76, 066203 (2007). 17. Gosak, M., Korosak, D. & Marhl, M. Optimal network configuration for maximal coherence resonance in excitable systems. Physical Review E 81, 056104 (2010). 18. Kaluza, P., Strege, C. & Meyer-Ortmanns, H. Noise as control parameter in networks of excitable media: role of the network topology. Physical Review E 82, 036104 (2010). 19. Sonnenschein, B., Zaks, M., Neiman, A. & Schimansky-Geier, L. Excitable elements controlled by noise and network structure. The European Physical Journal Special Topics 222, 2517 -- 2529 (2013). 20. Gollo, L. L., Kinouchi, O. & Copelli, M. Single-neuron criticality optimizes analog dendritic computation. Scientific Reports 3 (2013). 21. Kinouchi, O. & Copelli, M. Optimal dynamical range of excitable networks at criticality. Nature Physics 2, 348 -- 351 (2006). 22. Larremore, D. B., Shew, W. L. & Restrepo, J. G. Predicting criticality and dynamic range in complex networks: effects of topology. Physical Review Letters 106, 058101 (2011). 23. London, M. & Hausser, M. Dendritic computation. Annu. Rev. Neurosci. 28, 503 -- 532 (2005). 24. Stuart, G., Spruston, N. & Hausser, M. Dendrites (Oxford University Press, 2016). 25. Eagles, J. P. & Purple, R. L. Afferent fibers with multiple encoding sites. Brain Research 77, 187 -- 193 (1974). 26. Quick, D., Kennedy, W. & Poppele, R. Anatomical evidence for multiple sources of action potentials in the afferent fibers of muscle spindles. Neuroscience 5, 109 -- 115 (1980). 27. Banks, R. W., Barker, D. & Stacey, M. Form and distribution of sensory terminals in cat hindlimb muscle spindles. Philosophical Transactions of the Royal Society of London B: Biological Sciences 299, 329 -- 364 (1982). 12/15 28. Banks, R., Hulliger, M., Scheepstra, K. & Otten, E. Pacemaker activity in a sensory ending with multiple encoding sites: the cat muscle spindle primary ending. The Journal of Physiology 498, 177 -- 199 (1997). 29. Banks, R., Hulliger, M., Saed, H. & Stacey, M. A comparative analysis of the encapsulated end-organs of mammalian skeletal muscles and of their sensory nerve endings. Journal of Anatomy 214, 859 -- 887 (2009). 30. Besson, J. The neurobiology of pain. The Lancet 353, 1610 -- 1615 (1999). 31. Lesniak, D. R. et al. Computation identifies structural features that govern neuronal firing properties in slowly adapting touch receptors. Elife 3, e01488 (2014). 32. Walsh, C. M., Bautista, D. M. & Lumpkin, E. A. Mammalian touch catches up. Current Opinion in Neurobiology 34, 133 -- 139 (2015). 33. Lee, L.-Y. & Yu, J. Sensory nerves in lung and airways. Comprehensive Physiology (2014). 34. Kromer, J. A., Schimansky-Geier, L. & Neiman, A. B. Emergence and coherence of oscillations in star networks of stochastic excitable elements. Physical Review E 93, 042406 (2016). 35. Ermentrout, B. & Terman, D. H. Foundations of Mathematical Neuroscience (Springer Berlin, 2010). 36. Stemmler, M. A single spike suffices: the simplest form of stochastic resonance in model neurons. Network: Computation in Neural Systems 7, 687 -- 716 (1996). 37. Cover, T. M. & Thomas, J. A. Elements of Information Theory (John Wiley & Sons, 2012). 38. Kouvaris, N. E., Isele, T., Mikhailov, A. S. & Scholl, E. Propagation failure of excitation waves on trees and random networks. EPL (Europhysics Letters) 106, 68001 (2014). 39. Van Kampen, N. G. Elimination of fast variables. Physics Reports 124, 69 -- 160 (1985). 40. Da Fonseca, C. & Petronilho, J. Explicit inverses of some tridiagonal matrices. Linear Algebra and its Applications 325, 7 -- 21 (2001). 41. Seung, H. S. & Sompolinsky, H. Simple models for reading neuronal population codes. Proceedings of the National Academy of Sciences 90, 10749 -- 10753 (1993). 42. Brunel, N. & Nadal, J.-P. Mutual information, fisher information, and population coding. Neural Computation 10, 1731 -- 1757 (1998). 43. Pitkow, X., Liu, S., Angelaki, D. E., DeAngelis, G. C. & Pouget, A. How can single sensory neurons predict behavior? Neuron 87, 411 -- 423 (2015). 44. Greenwood, P. E. & Ward, L. M. Stochastic Neuron Models, vol. 1 (Springer, 2016). 45. Gammaitoni, L., Hanggi, P., Jung, P. & Marchesoni, F. Stochastic resonance. Reviews of Modern Physics 70, 223 (1998). 46. Pikovsky, A. S. & Kurths, J. Coherence resonance in a noise-driven excitable system. Physical Review Letters 78, 775 -- 778 (1997). 47. Pikovsky, A., Zaikin, A. & de La Casa, M. A. System size resonance in coupled noisy systems and in the ising model. Physical Review Letters 88, 050601 (2002). 48. Drmota, M. Random Trees: An Interplay Between Combinatorics and Probability (Springer Science & Business Media, 2009). 13/15 Supplementary material. Emergent stochastic oscillations and signal detection in regular tree networks of strongly coupled excitable elements Justus A. Kromer, Ali Khaledi-Nasab, Lutz Schimansky-Geier, Alexander B. Neiman Models for Nodes of Ranvier A Hodgkin-Huxley type model (HH) for a node of Ranvier contains only sodium and leak ionic currents. Thus, in eq.(2) of the main paper, the ionic current becomes Iion = INa + IL. For the sodium current we used the Hodgkin-Huxley (HH) type kinetics1, 2, INa = gNam3h(V − VNa), where gNa = 1100 mS/cm2 is the maximal value of the sodium conductance and VNa = 50 mV is the Na reversal potential. The gating activation and inactivation variables obey the dynamics m = a m(V )(1− m)− b m(V )m hk = a h(V )(1− h)− b h(V )h, with the following rate functions: a m(V ) = 1.314(V + 20.4)/[1− exp[−(V + 20.4)/10.3]], b m(V ) = −0.0608(V + 25.7)/[1− exp[(V + 25.7)/11]], a h(V ) = −0.068(V + 114)/[1− exp[(V + 114)/11]], b h(V ) = 2.52/[1 + exp[−(V + 31.8)/13.4]]. The leak current is IL = gL(Vk −VL) with gL = 20 mS/cm2 and VL = −80 mV. (24) (25) The Frankenhaeuser-Huxley (FH) model3 uses four ionic currents: sodium, potassium, persistent sodium and leak: Iion = INa + IK + Ip + IL. The currents are given by EF2 RT IX = PX IL = gL(V −VL), [X]0 − [X]ie EF RT 1− e EF RT , X = Na, K, p, (26) (27) where E = −70 +V; PX are the permeabilities of sodium (Na), potassium (K) and persistent (p) ionic currents: PNa = ¯PNahm2, PK = ¯PKn2, Pp = ¯Pp p2, where ¯PX are maximal values of permeabilities of corresponding channels. (27) [X]0 and [X]i are corresponding extracellular and intracellular ionic concentrations, of Na and K ions, respectively; for the persistent current, X = p, we have [X] ≡ [Na]. The gating variables follow, m = [a m(V )(1− m)− b m(V )m]s, h = [a h(V )(1− h)− b h(V )h]s, n = [a n(V )(1− n)− b n(V )n]s, p = [a p(V )(1− p)− b p(V )p]s, In Eq. (28) and the rate functions are: a m = 0.36(V − 22)/(1− exp[(22−V)/3]), b m = 0.4(13−V)/(1− exp[(V − 13)/20]), a h = 0.1(10 +V)/(exp[(V + 10)/6]− 1), b h = 4.5/(1 + exp[(45−V)/10]), a p = 0.006(V − 40)/(1− exp[(40−V)/10]), b p = 0.09(25 +V)/(exp[(V + 25)/20]− 1), a n = 0.02(V − 35)/(1− exp[(35−V)/10]), b n = 0.05(10−V)/(1− exp[(V − 10)/10]), where s is a scaling factor. The parameters for the model were taken from the original Frankenhaeuser-Huxley paper3 with three modifications to reduce the frequency of periodic spiking when a sufficiently-high constant current is injected: (i) the rate equations for the gating variables included a scale factor s = 0.3; (ii) the maximal permeability of potassium channels is reduced to ¯PK = 3.6 × 10−4 cm/sec; and (iii) the leak conductance was reduced to gL = 6 mS/cm2. Other parameters are the same as in the original FH model: [Na]0 = 114.5 mM, [Na]i = 13.74 mM, [K]0 = 2.5 mM, [K]i = 120 mM; ¯PNa = 8× 10−3 cm/sec; ¯Pp = 54× 10−5 cm/sec; VL = 0.026 mV. Constants R and F are the universal gas constant and the Faraday constant, and T = 293.15 K. 14/15 Inverse of Tridiagonal Toeplitz Matrix In order to calculate the effective current Imod eq. (20) and noise intensity Dmod eq. (21), we need to evaluate the inverse matrix B−1 of B. To this end, we apply results from Ref.4 in which the components of the inverse of a tridiagonal Toeplitz matrix are given in terms of two sequences {vi} and {ui} with i = 1,2, ...,G. If we apply their results to the matrix B, eq. (17), we obtain for its (i, j)th component B−1 i j = uMin(i,j)vMax(i,j)( di d j 1 if i < j else . (29) The sequences {ui} and {vi} are only defined up to a multiplicative constant and it is convenient to set u1 = 1. Then the sequences can be calculated from v1 = 1 d1 , vk = − d dk vk−1, k = 2,3, ..,G dG = − (d + 1), di = − (d + 1)− d di+1 , i = 1, ...,G− 2,G− 1 uG = 1 wGvG , uk = − d wk uk+1, k = 1, ....,G− 2,G− 1 with and with w1 = −(n1 + 1), wi = − (d + 1)− , i = 2,3, ..,G. d wi−1 Since we are interested in the dynamics of the first order node (g = 0) when only peripherals are subject to noisy currents, 1,G = vG. Applying we only need to evaluate a single component, (1,G), of B−1. For this component we find from eq. (29), B−1 the definition of vk, eq. (30), multiple times, we obtain vG = (−1)G−1 dG−1 G(cid:213) k=1 dk . Next, we evaluate the product for which one can show by using mathematical induction that j k=1 dk = (−1) j j k=0 dk − d d j+1 j−1 l=0 dk! , j < G. Finally, multiplication by dG and applying eq. (30) yields G(cid:213) k=1 dk = (−1)G G(cid:229) l=0 dl = (−1)GN. This yields for the (1,G) component of B−1 B−1 1G = − dG−1 N . (30) (31) (32) (33) Using this in eqs. (19) yields the effective current and noise intensity, respectively. References 1. McIntyre, C. C., Richardson, A. G. & Grill, W. M. Modeling the excitability of mammalian nerve fibers: influence of afterpotentials on the recovery cycle. Journal of Neurophysiology 87, 995 -- 1006 (2002). 2. Hodgkin, A. L. & Huxley, A. F. A quantitative description of membrane current and its application to conduction and excitation in nerve. The Journal of Physiology 117, 500 (1952). 3. Frankenhaeuser, B. & Huxley, A. The action potential in the myelinated nerve fibre of xenopus laevis as computed on the basis of voltage clamp data. The Journal of Physiology 171, 302 (1964). 4. Da Fonseca, C. & Petronilho, J. Explicit inverses of some tridiagonal matrices. Linear Algebra and its Applications 325, 7 -- 21 (2001). 15/15 (cid:213) (cid:229) (cid:229)
1609.03951
1
1609
2016-09-13T17:44:36
Architecture and Co-Evolution of Allosteric Materials
[ "physics.bio-ph", "cond-mat.mtrl-sci", "cond-mat.soft" ]
We introduce a numerical scheme to evolve functional materials that can accomplish a specified mechanical task. In this scheme, the number of solutions, their spatial architectures and the correlations among them can be computed. As an example, we consider an "allosteric" task, which requires the material to respond specifically to a stimulus at a distant active site. We find that functioning materials evolve a less-constrained trumpet-shaped region connecting the stimulus and active sites and that the amplitude of the elastic response varies non-monotonically along the trumpet. As previously shown for some proteins, we find that correlations appearing during evolution alone are sufficient to identify key aspects of this design. Finally, we show that the success of this architecture stems from the emergence of soft edge modes recently found to appear near the surface of marginally connected materials. Overall, our in silico evolution experiment offers a new window to study the relationship between structure, function, and correlations emerging during evolution.
physics.bio-ph
physics
r a i i "allosteryarxiv" -- 2018/11/13 -- 21:07 -- page 1 -- #1 Architecture and Co-Evolution of Allosteric Materials Le Yan ∗, Riccardo Ravasio †, Carolina Brito ‡, Matthieu Wyart † ∗Kavli Institute for Theoretical Physics, University of California, Santa Barbara, CA 93106, USA,†Institute of Physics, EPFL, CH-1015 Lausanne, Switzerland, and ‡Instituto de F´ısica, Universidade Federal do Rio Grande, do Sul CP 15051, 91501-970 Porto Alegre RS, Brazil Submitted to Proceedings of the National Academy of Sciences of the United States of America What are the correlations among them, and are these corre- lations sufficient by themselves to identify key aspects of the architecture, as proposed for real proteins? In this work, we answer these questions by introducing a model of elastic networks that can evolve according to some fitness function F , which depends on the response of the ma- terial to a well-defined stimulus. Our approach allows for a considerable freedom in the choice of the fitness function. As an illustration, we impose here that a displacement of four nodes on one side of the material (the "stimulus") elicits a given displacement of identical amplitude but different direc- tion on four target nodes on the other side of the system. A key advantage of our scheme is that our algorithm uniformly samples the fitness landscape (we use a Monte Carlo algorithm that turns out to equilibrate rapidly), which allows us to count the number of solutions and compute the entropy S(F ), as well as to guarantee that the solutions generated are the typ- ical (most numerous) ones. The quality of the solutions can be monitored by an "evolution temperature" Te that controls the fitness of the solutions probed. Our central findings are that: (i) there exists a transition temperature below which high-quality solutions appear and above which solutions are poor. (ii) High-quality solutions share a specific design. They present a trumpet-shaped region where the material is less constrained, which end by a marginally-constrained region in the vicinity of the target. (iii) The response amplitude varies non-monotonically between the stimulus and active sites. (iv) We rationalize this design based on a recent theory of edge modes in marginally connected disordered media [18]. (v) We show that co-evolution -- the correlations in the structures of the family of solutions -- alone is sufficient to identify the Significance In allosteric proteins, binding a ligand affects activity at a dis- tant site. The physical principle allowing for such an action at a distance are not well understood. Here we introduce a numeri- cal scheme to evolve allosteric materials in which the number of solutions, their spatial architectures and the correlations among them can be computed. We show that allostery in these mate- rials uses recently discovered elastic edge modes near the active site to transmit information, and that correlations generated during evolution alone can reveal key aspects of this architec- ture. Reserved for Publication Footnotes We introduce a numerical scheme to evolve functional materials that can accomplish a specified mechanical task. In this scheme, the number of solutions, their spatial architectures and the correlations among them can be computed. As an example, we consider an "al- losteric" task, which requires the material to respond specifically to a stimulus at a distant active site. We find that functioning ma- terials evolve a less-constrained trumpet-shaped region connecting the stimulus and active sites, and that the amplitude of the elastic response varies non-monotonically along the trumpet. As previously shown for some proteins, we find that correlations appearing during evolution alone are sufficient to identify key aspects of this design. Finally, we show that the success of this architecture stems from the emergence of soft edge modes recently found to appear near the surface of marginally connected materials. Overall, our in silico evolution experiment offers a new window to study the relationship between structure, function and correlations emerging during evolu- tion. Evolution Disordered materials Proteins Proteins are long polymers that can fold in a reproducible way and achieve a specific function. Often, the activity of the main functional site depends on the binding of an effector on a distant site [1]. Such an allosteric behavior can occur over large distances, such as 20 residues or more [2], and of- ten involves only a sparse subset of residues in the protein [2, 3]. Allosteric regulation offers an appealing target for drug design [4], and there is a considerable interest in predicting allosteric pathways [5, 6]. One central difficulty is that the physical mechanisms allowing such an action-at-a-distance re- main elusive. In some cases, allostery can be understood as the modulation of a hinge connecting two extended rigid parts of the protein [7, 8], but often the displacement field induced by the binding of the effector cannot be described in these terms [9, 10, 3]. Another route, statistical coupling analysis [11], considers correlations within sequences of proteins of the same family to infer allosteric pathways [3, 6]. The generality of this elegant approach is however debated [12]. From a physical viewpoint, specific response at a distance is surprising. The structure of proteins is similar to randomly packed spheres [13]. Generically, the response of such systems is non-specific and decays rapidly in space (in a manner similar to a continuum medium) at distances larger than the particle size. This is true except close to a critical point where the number of constraints coming from strongly interacting parti- cles is just sufficient to match the number of degrees of freedom of the particles [14]. There, the elastic response becomes het- erogeneous on all scales [15, 16]. This point is illustrated in Fig. 1.A showing the rapidly-decaying response of a random spring network to a stimulus. However, as shown in Fig. 1.B (and independently found in [17] using a different algorithm), springs can be moved so that the response extends further and specifically matches a target response on the other side of the system. This observation raises various questions, including: (i) Which network architectures allow for such allosteric re- sponse? (ii) Why are such architectures working from a phys- ical viewpoint? (iii) What is the number of solutions? (iv) www.pnas.org -- -- PNAS Issue Date Volume Issue Number 1 -- 10 i i i i i i "allosteryarxiv" -- 2018/11/13 -- 21:07 -- page 2 -- #2 i i trumpet structure. Finally, this detailed characterization of the solutions also points to some of their limitations in using them in thermal environments. We discuss how the fitness function can be changed to alleviate such problems. number of links. We define the average coordination number z as 2Ns/N . Marginally connected networks corresponds to zc = 4 in two dimensions [25], and we denote δz = z − zc. For a given configuration σ(cid:105), we consider the response to a ligand binding event, which we model as an imposed displace- ment field δRE(cid:105) on a set of 4 adjacent nodes (the "allosteric" site) located on one free boundary of the system, as illus- trated in purple in Fig. 1. After relaxing the elastic energy, such a stimulus will generate a displacement field δR(σ)r(cid:105) in the entire system; a fast numerical calculation of this re- sponse is formulated in Methods. Here, we focus on studying networks for which the response generates a desired target dis- placement δRT (cid:105) of identical amplitude but different direction (illustrated in cyan in Fig. 1) on an "active" site, which we also choose to consist of nT = 4 nodes on the other side of the system. Because physically only the strain (and not the absolute displacement) at the active site can affect catalytic properties, the target displacement is defined modulo a global translation and rotation U(cid:105). To rank networks in term of their allosteric ability, we define a fitness function F and a cost E: (cid:115)(cid:88) i∈T F (σ) ≡ −E(σ) ≡ − minU(cid:105) (δR(σ)r i − δRT i − Ui)2, [ 1 ] Fig. 1. Illustration of the model. A network is defined by the location of strong springs (in red), connecting the adjacent nodes of a distorted triangular lattice. A net- work is fit if its response field (black arrows) to an imposed stimulus (purple arrows) reproduces a target displacement (cyan arrows) at a distant site, consisting here of four nodes. In (A), a random configuration (Te = ∞) performs poorly. In (B), the system has evolved by moving strong springs, and performs almost perfectly. Here Te = 0.01, L = 12 and z = 5.0. Description of the evolution model Scalar models, where the response of a node is described by a scalar instead of a vector, have been introduced to study co- evolution and allostery [19, 20]. Although these models can capture the rigid motion of a part of the system, they cannot address the propagation of more complex mechanical informa- tion, such as that illustrated in Fig. 1. Instead, we use elastic networks, which have been used extensively to describe the vibrational dynamics of proteins [21, 22]. Specifically, we consider on-lattice models previously used to describe covalent glasses [23, 24]. N = L2 nodes are located on a triangular lattice (slightly distorted to avoid straight lines of nodes, see discussion in S.I.). Ns strong springs of stiff- ness ks connect some of the adjacent nodes, modeling strong interactions such as peptide bonds, disulfide bridges or other electrostatic interactions. We declare that σα = 1 if a strong spring is present in the link α (as represented in red in Fig. 1), and σα = 0 otherwise. To ensure that the elastic response is always uniquely defined, each node also interacts with all its next nearest neighbors via weak springs of stiffness kw = 10−4, mimicking weaker interactions. Thus the network is entirely described by a connection vector σ(cid:105) whose dimension is the where i label the nodes, and T the set of nodes belonging to the active site. Thus F = 0 corresponds to a perfect allosteric response. Below we restrict the networks further by impos- ing that all adjacent active nodes are connected, and choose a target displacement that does not stretch these bonds. The minimization of Eq.(1) can be readily performed and the fit- ness can be written directly in terms of δRT (cid:105) and δR(σ)r(cid:105), as discussed in Methods. Fig. 2. A. Average fitness (cid:104)F(cid:105) versus evolution temperature Te for various system sizes where z = 5.0. A steep change in fitness is seen near Tc ≈ 0.09 for relatively large systems. B. Specific heat c versus temperature Te. The max- imal specific heat increases with the system size L, suggesting the existence of a thermodynamic transition at Tc. C. Fitness averaged over local maxima (cid:104)F(cid:105)Te=0 versus coordination number z in log-linear scale. The black line shows the fitness if no mechanical response is present at the active sites. D. Entropy density S/N versus temperature Te. The entropy jump near Tc indicates the number of degrees of freedom that must be tuned to achieve the desired response. 2 www.pnas.org -- -- Footline Author i i i i Target(A)(B)StimulustemperatureTe00.10.20.3averagefitnesshFi(Te)-2-1.5-1-0.50L=4L=6L=8L=10L=12coordinationnumberz−zc-101maximumfitnesshFi(z)-101-100-10-1-10-2-10-3temperatureTe00.10.20.3specificheatc(Te)010203040L=4L=6L=8L=10L=12temperatureTe00.10.20.3entropyS/N00.20.40.60.811.2(D)(B)(C)(A)z=5.0 i i "allosteryarxiv" -- 2018/11/13 -- 21:07 -- page 3 -- #3 i i Henceforth we consider two dimensional networks with pe- riodic boundaries in the direction transverse to the direction between the allosteric and active sites (this corresponds to a cylindrical geometry). As is the case for many aspects of the microscopic elasticity of amorphous materials [14], we expect our results to hold independently of the spatial dimension. works. Z =(cid:80) Numerical solutions To evolve networks we use a Metropolis algorithm (see Meth- ods) at some "evolution temperature" Te, where strong springs can swap from an occupied to an unoccupied link, leaving the average coordination z constant. More precisely, Te is the variable conjugate to the fitness, so that once our al- gorithm reaches equilibrium (which it does in the range of Te values presented here), the probability P (σ) of finding a configuration σ(cid:105) reads P (σ) = exp(F (σ)/Te)/Z where σ exp(F (σ)/Te). Te = ∞ corresponds to random net- Thus as we lower the temperature, we probe fitter and fitter networks as illustrated in Fig. 1. This point is systematically studied in Fig. 2.A showing (cid:104)F(cid:105)(Te) for a given coordina- tion, where the ensemble average is made on both Monte-Carlo steps and different realizations. For Te > 0, we find that this average does not depend on the time of the simulation if it is long enough, indicating that an equilibrium was reached. As Te decreases, we observe a transition from low to high fitness, which appears to become sharper and sharper as N increases. This suggests a thermodynamic transition, as we evidence fur- ther by considering the specific heat c(Te) = d(cid:104)E(cid:105)(Te) which displays a more and more pronounced peak as L increases, as shown in Fig. 2.B. This result is a signature of an underlying collective phenomenon, indicating that achieving an allosteric function is a collective process. Below the transition, we find that the networks perform well. Their performance can be quantified by (cid:104)F(cid:105)Te=0, an ensemble average of local maxima in the fitness landscape, as reported in Fig. 2.C. These structures result from a pure gradient ascent in the fitness landscape. We find that in the range of coordination we probed, the cost decreases by at least 200 folds with respect to random networks, i.e. the response converges very precisely toward the desired one. Thus, the systems does not get stuck in local maxima of poor quality in the fitness landscape. dTe , and is shown in Fig. Finally, we can quantify the number of allosteric networks. It follows eS(Te), where S(Te) is the entropy. It satisfies dS = c(Te) dTe 2.D. For example, Te at Te = 0.05 where networks perform very well, we find that their number is very large -- exponential to the system size: eS(Te) ≈ 1053 for L = 12, but the probability pA to obtain such a network by chance is also exponentially low: pA = eS(Te)−S(∞) ≈ 10−10 for L = 12. Architecture of allosteric networks Hypostatic networks with δz < 0 are extremely floppy. It may be a interesting case to study intrinsically disordered proteins [26], but for folded proteins considering δz > 0 is more real- istic. Henceforth we focus on that case and choose z = 5. In S.I. our results are presented for the floppy case δz < 0. Which architecture allows for such a long-distance, specific response? A systematic design is revealed by averaging the occupancy of various solutions which our algorithm generates, as shown in Fig. 3.A. At high temperature, the structures are essentially random and not functioning. At low temperature, a trumpet-shaped region appears that connects the allosteric and active sites. Specific features are that: • Inside the trumpet, the mean occupancy is lower than the mean, but there are no floppy modes (i.e. modes that do not deform the strong springs). • The mean occupancy or coordination decreases monotoni- • The mouthpiece of the trumpet is surrounded by two more • The coordination number is close to its critical value, called rigid regions, that appear in dark in Fig. 3.A. cally from the allosteric to the active site. isostatic, in the vicinity of the active site (see below). The trumpet-like architecture is robust: it remains qualita- tively unchanged as the mean coordination number is varied, as long as δz > 0. For δz < 0 however, a trumpet still exists (see SI), but it is inverted: it is more coordinated than the rest of the system. Fig. 3. A. Map of the mean coordination number and B. Spatial distribution of the average response magnitude for configurations equilibrated at Te = 0.30 (left) and Te = 0.05 (right). In the functioning networks (right), a trumpet connect- ing the allosteric and active sites appear in A, and the response to stimulus varies non-monotonically inside the trumpet in B. i δRr i2/(cid:80) Next we study how such trumpets shape the response to a binding event, by considering the mean-squared mag- nitude of the normalized response at different nodes i, i2(cid:105) as shown in Fig. 3. For random net- (cid:104)δRr works, unsurprisingly the response is large only close to the stimulus site. However, the response of fit networks display a striking feature: it varies non-monotonically between the allosteric and the active site. It almost vanishes in the bulk of the material, but reappears near the active site where it is the strongest. Physical processes underlying allostery The observation that fit networks develop a less-constrained region connecting the stimulus to the active site is not very surprising, since the elastic point response can remain hetero- geneous on longer length scales in that case [16]. This ar- gument does not explain however the strong asymmetry of the trumpet, more coordinated near the stimulus and nearly marginally connected near the active site. We now argue that Footline Author PNAS Issue Date Volume Issue Number 3 i i i i 44.555.5Te=0:05Te=0:30(A)00.020.040.06(B)Te=0:30Te=0:05 i i "allosteryarxiv" -- 2018/11/13 -- 21:07 -- page 4 -- #4 i i Fig. 4. (A) Illustration of the mechanism responsible for allostery in our artificial networks.They display a nearly isostatic region in the vicinity of the active site, surrounded by a better connected material. When the ligand binds, it induces an effective shape change at the allosteric site. This mechanical signal transmits and decays through the well-connected body of the material. It is then amplified exponentially fast in the isostatic region near the active site, leading to a large strain. (B) System made of two elastic networks with coordination z = 5.0 (red) and z = 4.0 (green). Its response (black) to a perturbation (purple) demonstrates that a marginal network embedded in a more connected one can amplify the response near its free boundary. (C) Spatial distribution of the probability that a node is in an isostatic region connected to the active sites at Te = 0.3 and Te = 0.05 for z = 5.0. this design is selected for because it prevents the decay in the amplitude of the signal one expects in traditional materials. Recent works have shown that marginally connected crys- tals can display edge modes, leading to exponentially growing response when displacement are imposed at the boundary of the system [27, 28, 29]. It was very recently shown that such "explosive" modes must be present in disordered marginally connected materials as well [18]. Such systems, if sufficiently constrained at some of their boundaries, can act as a lever which can amplify complex motions exponentially toward free boundaries [18]. We argue that our allosteric networks are built along this principle. As sketched in Fig. 4.A., their structure is approx- imately that of a well-connected elastic material surrounding a marginally connected (called "isostatic") network near the active site. If a stimulus is imposed on an allosteric site, the response will decay with distance in the well-connected region, leading to imposed displacements of small amplitude on the boundary of the isostatic region. As noted above, "squeez- ing" such systems leads to an explosive response in the di- rection of free boundaries, allowing the response to reach the desired amplitude on the target nodes. This explanation cap- tures both the observation that allosteric networks are nearly isostatic near the active site, and that their response varies non-monotonically in space. To test this proposal we build an artificial network, by em- bedding a random isostatic network with z = 4 to a better coordinated random network with z = 5 (the details of the construction are explained in Methods). As shown in Fig 4.B, the response to an imposed dipole on the open boundary of the well-connected network decays, but eventually grows rapidly toward the boundary of the isostatic region, as predicted. As discussed in Methods, we have developed an algorithm to recognize a nearly isostatic region that contains the target nodes in our evolved networks. In Fig. 4.C we show the prob- ability map that a node belongs to such an isostatic network, obtained by averaging on many realizations. Fit allosteric net- works indeed show a robust isostatic region attached to the active sites, which is absent for random networks. Σα ≡ (cid:104)σα(cid:105) ln + (1 − (cid:104)σα(cid:105)) ln (cid:104)σα(cid:105) ¯σ 1 − (cid:104)σα(cid:105) 1 − ¯σ Conservation and co-evolution In our model, the "sequence" of a network corresponds to the vector of zeros and ones σ(cid:105) that defines a structure. This is analogous to the sequence of amino acids that defines a protein. Using our Monte-Carlo at some low Te, we can gen- erate a family of sequences associated with networks of high fitness. If only such a family could be observed (and assuming no knowledge on the task being performed nor on the spatial organization of the networks), would it be possible to infer which region of the system matters for function? There is evi- dence that such inference is useful for some protein families, if enough sequences are available [6, 3]. We show that it is also the case in our model. Key aspects of the design are more likely to stay conserved in evolution. Here we define conservation in each link α as [30]: , [ 2 ] where (cid:104)σα(cid:105) is the ensemble average of the occupancy of link α and ¯σ = Ns/(3N −2L) is the mean occupancy of the links. Σα is a measure of the predicability of the occupancy of the link α: it is zero when the link occupancy is random, and maximum when the link is always empty or always occupied. The con- servation map of allosteric networks is shown in Fig. 5.A. We can distinguish the trumpet pattern, but most strikingly, the neighborhood of the active site is very conserved. This obser- vation supports that specificity of the response is essentially controlled by the geometry of the network near the active site. Next, we test if an analysis of co-evolution alone reveals important features of function and structure. A more quanti- tative discussion will be presented elsewhere, here we highlight the main results. We define the correlation matrix C between the links α, β as: Cαβ = (cid:104)σασβ(cid:105) − (cid:104)σα(cid:105)(cid:104)σβ(cid:105), [ 3 ] where the (cid:104)•(cid:105) is again the ensemble average over the solutions found by the Monte Carlo algorithm. We then compute the eigenvalues λ1 > λ2 > ... > λNs of the matrix C. Fig. 5.B compares the spectrum of eigenvalues of a high temperature 4 www.pnas.org -- -- Footline Author i i i i active siteligandallosteric sitemore rigidnearlyisostatic(A) i i "allosteryarxiv" -- 2018/11/13 -- 21:07 -- page 5 -- #5 i i NΓ(cid:88) (essentially random) network with that of allosteric networks obtained at small Te. In the latter case, some eigenvalues are much larger than the continuum spectrum, itself much more spread than in the random case. To select out a "sector" [6] of links that co-evolve, we: (i) pick up the NΓ = 10 eigenvectors ψγ(cid:105) with highest eigenval- ues; (ii) we include a given link α in the sector if for at least one α >  = 0.05. Links selected in this pro- of this ten modes, ψγ cedure are shown in Fig. 5.C. This procedure selects precisely the links that belong to the trumpet, supporting the idea that co-evolution alone can uncover key functional aspects [6, 3]. For completeness, we define a correlation matrix recon- structed from its 10 top eigenvectors: Cαβ = λγψγ(cid:105)(cid:104)ψγ. [ 4 ] γ=1 C is shown in Fig. 5.D after re-ordering links in terms of the strength of their components in the top ten modes, clearly showing a sector of links where correlations are strong in am- plitude (but vary in sign). 5. Fig. (A) Spatial distribution of conservation, as defined in the text, for Te = 0.05 and z = 5.0. (B) Spectrum of eigenvalues ρ(λ) of C for the high temperature case (Te = 0.30) in black and low temperature (Te = 0.05) in red. The arrow indicates which eigenvalues are used to identify the springs shown (C) Springs selected using the procedure explained in the text. (D) C is in (C). built using the same parameters as in (C). C presents a clear separation in a region where the correlations are stronger, which corresponds to the trumpet shown in (C). All these figures are made using L = 12 and z = 5. Conclusion We have introduced a scheme to discover materials that ac- complish a specified task. It allows us to characterize the architecture of the solutions, their entropy and how correlated they are. We illustrated this approach using a specific al- losteric task, where a strain imposed on an allosteric site must lead to a given strain on a distant active site. The architec- tures we obtain are highly anisotropic. Our analysis revealed that the physical mechanisms that enable allostery include the recently discovered presence of soft edge modes in weakly- connected elastic materials [18]. It would be very interesting to test if some proteins have evolved to exploit such surface effects to operate. In addition, our analysis supports the no- tion [6, 3] that analyzing the correlation of the "sequences" defining structure can indeed reveal compact regions central for function. The detailed study of the architectures we found also reveals some of their limitations. Real proteins have additional con- straints than those we have considered: among others, they are made of a chain that folds and remains relatively stable despite thermal noise. Our asymmetric structures are quite soft near the active site: as documented in S.I, the thermally- induced motion would be about four times larger there than in the other nodes also located at the system surface, which may not be desirable. Note however that such features will im- prove if alternative fitness functions are considered, which our approach allows for. This could be implemented by explicitly penalizing thermal motion at the active site. An intriguing extension of our work is to reason in terms of energy, instead of displacement, and maximize the cooperativity between two distant sites. Denoting by E1 and E2 the mechanical ener- gies associated with a binding event in some site 1 and site 2 respectively, and E12 the energy of binding both, we can consider F = E1 + E2 − E12. Fitness can be large only if the two sites are strongly coupled together elastically, which from the symmetry of the fitness presumably corresponds to more symmetric architectures than those discovered here. Methods Computing the linear response to an imposed displacement. The linear response to an external force field F(cid:105) reads: F(cid:105) = MδR(cid:105) [ 5 ] where the stiffness matrix M depends only on connection σ(cid:105) and the link directions. To impose the stimulus at the allosteric nodes δRE(cid:105), forces must be applied on these nodes. All other nodes adapt to a new mechanical equilibrium with no net forces on them, and follow a displacement δR(σ)r(cid:105). Thus Eq.[5] becomes: (cid:18) δRE(cid:105) δR(σ)r(cid:105) (cid:19) (cid:18) δRE(cid:105) . 0(cid:105) = Q−1M (cid:19) (cid:18) FE(cid:105) 0(cid:105) FE(cid:105) δR(σ)r(cid:105) Qij = = M (cid:19) (cid:26) δij Mij ifj ∈ E ifj (cid:54)∈ E (cid:19) [ 6 ] [ 7 ] [ 8 ] which leads to:(cid:18) with (cid:115)(cid:88) Computing the fitness. Minimizing Eq.(1) with respect to the global translation and rotation leads to: i − δRT i )2 −(cid:88) i+1 − δRT i+1) i )(δRr (δRr i − δRT (δRr i∈T i∈T where i + 1 = min(T ) if i = max(T ). Metropolis algorithm. Starting from a configuration σ(cid:105), we consider the move toward a new configuration σ(cid:48)(cid:105) that differs only by the motion of a strong spring. The move is accepted with the probability: [ 9 ] (cid:18) F (σ(cid:48)(cid:105)) − F (σ(cid:105)) (cid:19) P (σ(cid:105) → σ (cid:48)(cid:105)) = min[1, exp Te ], [ 10 ] which satisfies detailed balance. This finishes one Monte Carlo step. The next step starts from the configuration just sampled. Footline Author PNAS Issue Date Volume Issue Number 5 i i i i 00.20.40.6(A)(C)-0.01-0.00500.0050.01eigenvalue600.20.40.60.8spectrum;(6)100101102(D)(B) i i "allosteryarxiv" -- 2018/11/13 -- 21:07 -- page 6 -- #6 i i For each coordination number z and each evolution temper- ature Te, we sample 20 Monte Carlo sampling series with 105 Monte Carlo steps in each, and do not consider the first half of these time series (which is sufficient to eliminate transient effects). Our results are thus averaged over 106 configurations. Mosaic network. To construct the mosaic network of Fig. 4B, we first generate a highly coordinated random network by an- nealing N bidisperse soft discs with radius ratio 1.4 : 1 to mechanical equilibrium. We then produce networks with dif- ferent coordination numbers by removing springs between the most connected neighbors one by one [15, 31]. This procedure ensures that the spatial fluctuations of coordination are low. We then cut the periodic boundaries in two networks, with z = 4.0 with z = 5.0 respectively, obtained from the same packing (and thus presenting nodes at the same locations). The network with z = 5.0 is duplicated three times. We com- plete a four square made of these patches, and connect nodes through the patch boundaries. We apply a periodic boundary condition in the horizontal direction. Algorithm to detect the isostatic region around the active site. The pebble game algorithm [32] used to identify indepen- dent rigid clusters cannot reveal a nearly isostatic network embedded in an over-constrained region. Here we introduce an algorithm to do that. The basic idea is to search for a con- nected set of nodes including the active site where Maxwell's rigidity condition [25] is satisfied to a desired level of accuracy. Specifically, we: (i) Initialize the searching subnetwork with the target nodes. The subnetwork includes the nodes inside, the internal con- nections between these nodes and the external connections to the nodes outside of the subnetwork. (ii) Count the degrees of freedom Nf and the number of con- straints Nc in the subnetwork. Nf = dn0, where n0 is the number of nodes inside the subnetwork. Nc = nI +qnE, where nI and nE are the numbers of internal and external connec- tions. The parameter q ∈ [0, 1]. (iii) If Nf > Nc, compute N(cid:48) c for each external node linked by an external connection when including it into the subnetwork. (iii') Otherwise, stop. (iv) Update the subnetwork by including the nearest node with minimal N(cid:48) (v) Check and include the nodes all of whose neighbors are in the subnetwork. Go to step (iii). f and set Nf = N(cid:48) f and N(cid:48) c − N(cid:48) f Nc = N(cid:48) c. The algorithm identifies a compact region depending on the parameter q that counts the external connections. In the re- sults presented in Fig. 4.C, we have chosen q = 0.5 (q = 1 corresponds to the pebble game). We discuss the physical meaning of different q and show that q = 0.5 is an appropriate choice in SI. ACKNOWLEDGMENTS. We thank B. Bialek, J-P Bouchaud, P. De Los Rios, E. DeGiuli, D. Malinverni, R. Monasson, O. Rivoire and S. Zamuner for discussions. L.Y. was supported in part by the National Science Foundation under Grant No. NSF PHY11-25915. M.W. thanks the Swiss National Science Foundation for support un- der Grant No. 200021-165509 and the Simons Foundation Grant (#454953 Matthieu Wyart). This material is based upon work performed using computational resources supported by the "Center for Scientific Computing at UCSB" and NSF Grant CNS- 0960316. 1. Jean-Pierre Changeux and Stuart J Edelstein. Allosteric mechanisms of signal trans- 20. Tsvi Tlusty, Albert Libchaber, and Jean-Pierre Eckmann. Physical model of the duction. Science, 308(5727):1424 -- 1428, 2005. 2. Michael D Daily and Jeffrey J Gray. Local motions in a benchmark of allosteric proteins. Proteins: Structure, function, and bioinformatics, 67(2):385 -- 399, 2007. 3. Richard N McLaughlin Jr, Frank J Poelwijk, Arjun Raman, Walraj S Gosal, and Rama Ranganathan. The spatial architecture of protein function and adaptation. Nature, 491(7422):138 -- 142, 2012. 4. Ruth Nussinov and Chung-Jung Tsai. Allostery in disease and in drug discovery. Cell, 153(2):293 -- 305, 2013. 5. Benjamin RC Amor, Michael T Schaub, Sophia N Yaliraki, and Mauricio Barahona. Prediction of allosteric sites and mediating interactions through bond-to-bond propen- sities. Nature Communications, 7, 2016. 6. Najeeb Halabi, Olivier Rivoire, Stanislas Leibler, and Rama Ranganathan. Protein sec- tors: evolutionary units of three-dimensional structure. Cell, 138(4):774 -- 786, 2009. 7. MF Perutz. Stereochemistry of cooperative effects in haemoglobin: Haem -- haem in- teraction and the problem of allostery. Nature, 228:726 -- 734, 1970. 8. Mark Gerstein, Arthur M Lesk, and Cyrus Chothia. Structural mechanisms for domain movements in proteins. Biochemistry, 33(22):6739 -- 6749, 1994. 9. Nina M Goodey and Stephen J Benkovic. Allosteric regulation and catalysis emerge via a common route. Nat Chem Biol, 4(8):474 -- 482, 2008. 10. PS Gandhi, Z Chen, F Scott Mathews, and E Di Cera. Structural identification of the pathway of long-range communication in an allosteric enzyme. Proceedings of the National Academy of Sciences, 105(6):1832 -- 1837, 2008. 11. Steve W Lockless and Rama Ranganathan. Evolutionarily conserved pathways of en- ergetic connectivity in protein families. Science, 286(5438):295 -- 299, 1999. 12. Tiberiu Te¸sileanu, Lucy J Colwell, and Stanislas Leibler. Protein sectors: statistical coupling analysis versus conservation. PLoS Comput Biol, 11(2):e1004091, 2015. 13. Jie Liang and Ken A Dill. Are proteins well-packed? Biophysical Journal, 81(2):751 -- 766, 2001. 14. Andrea J. Liu, Sidney R. Nagel, Wim van Saarloos, and Matthieu Wyart. The jamming scenario: an introduction and outlook. Oxford University Press, Oxford, 2010. 15. Gustavo During, Edan Lerner, and Matthieu Wyart. Phonon gap and localization lengths in floppy materials. Soft Matter, 9(1):146 -- 154, 2013. 16. Edan Lerner, Eric DeGiuli, Gustavo During, and Matthieu Wyart. Breakdown of con- tinuum elasticity in amorphous solids. Soft Matter, 10:5085 -- 5092, 2014. 17. Jason W Rocks, Nidhi Pashine, Irmgard Bischofberger, Carl P Goodrich, Andrea J Liu, and Sidney R Nagel. Designing allostery-inspired response in mechanical networks. arXiv preprint arXiv:1607.08562, 2016. 18. Le Yan, Jean-Philippe Bouchaud, and Matthieu Wyart. Edge mode amplification in disordered elastic networks. arXiv preprint arXiv:1608.07222, 2016. 19. Mathieu Hemery and Olivier Rivoire. Evolution of sparsity and modularity in a model of protein allostery. Physical Review E, 91(4):042704, 2015. sequence-to-function map of proteins. arXiv preprint arXiv:1608.03145, 2016. 21. AR Atilgan, SR Durell, RL Jernigan, MC Demirel, O Keskin, and I Bahar. Anisotropy of fluctuation dynamics of proteins with an elastic network model. Biophysical journal, 80(1):505 -- 515, 2001. 22. Paolo De Los Rios, Fabio Cecconi, Anna Pretre, Giovanni Dietler, Olivier Michielin, Francesco Piazza, and Brice Juanico. Functional dynamics of pdz binding domains: a normal-mode analysis. Biophysical journal, 89(1):14 -- 21, 2005. 23. Le Yan and Matthieu Wyart. Evolution of covalent networks under cooling: contrasting the rigidity window and jamming scenarios. Physical review letters, 113(21):215504, 2014. 24. Le Yan and Matthieu Wyart. Adaptive elastic networks as models of supercooled liquids. Physical Review E, 92(2):022310, 2015. 25. J.C. Maxwell. On the calculation of the equilibrium and stiffness of frames. Philos. Mag., 27(5755):294 -- 299, 1864. 26. A Keith Dunker, Christopher J Oldfield, Jingwei Meng, Pedro Romero, Jack Y Yang, Jessica W Chen, Vladimir Vacic, Zoran Obradovic, and Vladimir N Uversky. The unfoldomics decade: an update on intrinsically disordered proteins. BMC genomics, 9(Suppl 2):S1, 2008. 27. CL Kane and TC Lubensky. Topological boundary modes in isostatic lattices. Nature Physics, 10(1):39 -- 45, 2014. 28. Bryan Gin-ge Chen, Nitin Upadhyaya, and Vincenzo Vitelli. Nonlinear conduction via solitons in a topological mechanical insulator. Proceedings of the National Academy of Sciences, 111(36):13004 -- 13009, 2014. 29. Daniel M Sussman, Olaf Stenull, and TC Lubensky. Topological boundary modes in jammed matter. arXiv preprint arXiv:1512.04480, 2015. 30. Thomas M. Cover and Joy A. Thomas. Elements of information theory (2. ed.). John Wiley & Sons, 2006. 31. Le Yan and Matthieu Wyart. On variational arguments for vibrational modes near jamming. arXiv preprint arXiv:1601.02141, 2016. 32. D. J. Jacobs and M. F. Thorpe. Generic rigidity percolation: The pebble game. Phys. Rev. Lett., 75:4051 -- 4054, Nov 1995. 33. KN Trueblood, H-B Burgi, H Burzlaff, JD Dunitz, CM Gramaccioli, HH Schulz, U Shmueli, and SC Abrahams. Atomic dispacement parameter nomenclature. report of a subcommittee on atomic displacement parameter nomenclature. Acta Crystallo- graphica Section A: Foundations of Crystallography, 52(5):770 -- 781, 1996. 6 www.pnas.org -- -- Footline Author i i i i i i "allosteryarxiv" -- 2018/11/13 -- 21:07 -- page 7 -- #7 i i Footline Author PNAS Issue Date Volume Issue Number 7 i i i i i i "allosteryarxiv" -- 2018/11/13 -- 21:07 -- page 8 -- #8 i i Supplementary Information (SI) A. Distortion of triangular lattice In our model, we introduce a slight distortion of the lattice to remove long straight lines that occur in a triangular lattice. Such straight lines are singular and lead to unphysical local- ized floppy modes orthogonal to them. One can remove them by imposing a random displacement on the nodes. Instead, we distort the lines without introducing frozen disorder. We group nodes in lattice by four, labeled as A B C D in Fig. S1. One group forms a cell of our distorted lattice. In each cell, node A stays in place, while nodes B, C, and D move by some distance δ: B along the direction perpendicular to BC, C along the direction perpendicular to CD, and D along the direction perpendicular to DB, as illustrated. We set δ to 0.2, where the straight lines are maximally reduced with this distortion. pet" shape structure similar to the rigid case, connecting the allosteric site and the active site, as appears in the average co- ordination map shown in Fig. S2.A. The coordination number in this trumpet structure is larger than in the rest of the mate- rial, allowing the signal to propagate. Similarly to the δz > 0 case, the mean coordination number decreases monotonically from the allosteric to the active site. However, the trumpet has the wider nearly-isostatic patch near the allosteric side. We also find that the amplitude of the displacement decreases monotonically from the allosteric to the active side in the bot- tom panel of Fig. S2.B. C. Thermal noise We quantify thermally-induced motion in a structure by the B-factor [33] (related to the Debye-Waller factor), B = 8π2(cid:104)u2(cid:105) [ S1 ] where u is the magnitude of particle displacement around its mean position, and (cid:104)•(cid:105) denotes thermal averaging. Fig. S1. Illustration of the distorted triangular lattice. Fig. S3. (right). Colorbar is in log scale, z = 5.0. Spatial distribution of B-factor for Te = 0.30 (left) and Te = 0.05 In an elastic network, the energy due to displacement field (cid:126)ui is, H = 1 2 (cid:88) i,j (cid:126)ui · Mij · (cid:126)uj [ S2 ] where M is the stiffness matrix defined in Eq.(5). The B- factor is directly related to the stiffness matrix M, as we now recall. By definition, the thermal average of the correlation of the displacement at temperature T is (cid:80) i,j (cid:126)ui·Mij·(cid:126)uj , [ S3 ] − 1 2T (cid:104)(cid:126)uk · (cid:126)ul(cid:105)(T ) = 1 Z(T ) d(cid:126)ui(cid:126)uk · (cid:126)ule (cid:90) (cid:89) (cid:104)(cid:126)ui · (cid:126)uj(cid:105) = T(cid:0)M−1(cid:1) i , ij a Gaussian integral. It can thus be carried out, [ S4 ] [ S5 ] Fig. S2. A. Spatial distribution of coordination number. B. Spatial distribution of the response magnitude to the excitation at allosteric sites. Left: Te = 0.3. Right: Te = 0.10. z = 3.0. and Bi = 8π2(cid:104)u2 i(cid:105) = 8π2T (M−1)ii. B. "Trumpet" structure in floppy networks In general, the displacement signal can only propagate a finite distance in rather homogeneous floppy networks [15]. This can be seen in the left panel of Fig. S2.B. Below the transition temperature (Tc = 0.12 for z = 3.0), we find that the network resolves this issue by generating a "trum- We compute B-factor with temperature T = 1. In Fig. S3, we show the B-factor map for the configurations at high and low evolution temperature. Thermal motions are stronger in the"trumpet" shape discussed in the main text, especially at the vicinity of the active site. There, the B- factors are about 20 times larger than in the other nodes at the boundary of the system, corresponding to an amplitude of motion about four times larger. 8 www.pnas.org -- -- Footline Author i i i i DBACδ2.533.544.5Te=0:30Te=0:05(A)00.050.10.15(B)Te=0:30Te=0:05102.5103 103.5104 Te=0:05Te=0:30 i i "allosteryarxiv" -- 2018/11/13 -- 21:07 -- page 9 -- #9 i i Fig. 4. belonging to z = 4.0 and z = 5.0 patches in the mosaic network. From left to right, external constraints are counted as q = 0, q = 0.5, q = 1. Isostatic region identified by the constraint-counting algorithm for the mosaic network. Identified region is shown by green nodes. Green and red lines are links D. Counting of the external constraints We have described an algorithm to detect the nearly isostatic region near the target nodes, in which there is an undeter- mined parameter q, the fraction of counting an external con- straint. q describes the boundary condition of the subnetwork: q = 0 for open boundary condition; q = 1 for fixed boundary condition; q = 0.5 for a boundary equivalent to the bulk. We find that q = 0.5 works better than q = 0 or q = 1 numeri- cally in identifying the nearly-isostatic region for the mosaic network, shown in Fig. 4. Footline Author PNAS Issue Date Volume Issue Number 9 i i i i q=0:q=1:q=0.5:
1705.08425
2
1705
2017-12-18T15:43:05
Pattern formation by curvature-inducing proteins on spherical membranes
[ "physics.bio-ph", "cond-mat.soft" ]
Spatial organisation is a hallmark of all living cells, and recreating it in model systems is a necessary step in the creation of synthetic cells. It is therefore of both fundamental and practical interest to better understand the basic mechanisms underlying spatial organisation in cells. In this work, we use a continuum model of membrane and protein dynamics to study the behaviour of curvature-inducing proteins on membranes of spherical shape, such as living cells or lipid vesicles. We show that the interplay between curvature energy, entropic forces, and the geometric constraints on the membrane can result in the formation of patterns of highly-curved/protein-rich and weakly-curved/protein-poor domains on the membrane. The spontaneous formation of such patterns can be triggered either by an increase in the average density of curvature-inducing proteins, or by a relaxation of the geometric constraints on the membrane imposed by the membrane tension or by the tethering of the membrane to a rigid cell wall or cortex. These parameters can also be tuned to select the size and number of the protein-rich domains that arise upon pattern formation. The very general mechanism presented here could be related to protein self-organisation in many biological processes, ranging from (proto)cell division to the formation of membrane rafts.
physics.bio-ph
physics
Pattern formation by curvature-inducing proteins on spherical membranes Jaime Agudo-Canalejo Theory & Bio-Systems Department, Max Planck Institute of Colloids and Interfaces, 14424 Potsdam, Germany Rudolf Peierls Centre for Theoretical Physics, University of Oxford, Oxford OX1 3NP, United Kingdom and Department of Chemistry, The Pennsylvania State University, University Park, Pennsylvania 16802, United States Ramin Golestanian Rudolf Peierls Centre for Theoretical Physics, University of Oxford, Oxford OX1 3NP, United Kingdom and Max Planck Institute for the Physics of Complex Systems, Nothnitzer Str. 38, D-01187 Dresden, Germany (Dated: September 19, 2018) 7 1 0 2 c e D 8 1 ] h p - o i b . s c i s y h p [ 2 v 5 2 4 8 0 . 5 0 7 1 : v i X r a 1 Abstract Spatial organisation is a hallmark of all living cells, and recreating it in model systems is a necessary step in the creation of synthetic cells. It is therefore of both fundamental and practical interest to better understand the basic mechanisms underlying spatial organisation in cells. In this work, we use a continuum model of membrane and protein dynamics to study the behaviour of curvature-inducing proteins on membranes of spherical shape, such as living cells or lipid vesicles. We show that the interplay between curvature energy, entropic forces, and the geometric constraints on the membrane can result in the formation of patterns of highly-curved/protein-rich and weakly- curved/protein-poor domains on the membrane. The spontaneous formation of such patterns can be triggered either by an increase in the average density of curvature-inducing proteins, or by a relaxation of the geometric constraints on the membrane imposed by the membrane tension or by the tethering of the membrane to a rigid cell wall or cortex. These parameters can also be tuned to select the size and number of the protein-rich domains that arise upon pattern formation. The very general mechanism presented here could be related to protein self-organisation in many biological processes, ranging from (proto)cell division to the formation of membrane rafts. 2 I. INTRODUCTION Spatial organisation into inhomogeneous patterns is an essential feature of living organ- isms, from the macroscale to the cellular level. In the later case, organisation of the plasma membrane and the cytoplasm into specialised domains is more commonly referred to as cell polarity. [1, 2] This spatial organisation of the cell is necessary in order to coordinate important processes such as cell division, differentiation, or directed cell migration. As early as in 1952, Turing realised [3] that very simple systems that are initially in a spatially homogeneous state can spontaneously self-organise into spatially inhomogeneous patterns. However, it is generally believed [1, 2] that the generation of polarity in cells is the result of a tightly-controlled orchestration involving complex signalling networks and active processes such as the reorganisation of the cellular cytoskeleton. Nevertheless, active systems such as the cytoskeleton have been shown to undergo simple pattern formation, [4] and there also exist cells for which polarisation is presumably not generated by the cytoskeleton. [5 -- 11] The underlying mechanisms in these systems are however not well understood. Very recently, [12] a system was identified in which cell polarisation appears to be con- trolled by a relatively simple pattern-formation mechanism. In the coccal bacterium Staphy- lococcus aureus, essential proteins involved in lipid metabolism were seen to distribute in inhomogeneous spatial patterns, that could be explained by a model that considers the dy- namics of curvature-inducing proteins on a spherical membrane. However, the model first introduced in Ref. 12 is very general, and we expect that it might be able to describe the formation of protein patterns on the surface of other types of cells, as well as in model sys- tems consisting of lipid vesicles and proteins. In this work, we will explore in full generality and detail the predictions of such a model. The basic idea behind the model is presented in figure 1. A closed, initially spherical membrane contains proteins that impose a spontaneous curvature Cp on the membrane (in general, the proteins might be attached to the membrane from the cytoplasmic or the exoplasmic sides, or they might be transmembrane proteins embedded in the membrane). [13, 14] If the proteins did not induce any curvature, a random, homogeneous distribution of proteins would be favoured by thermal fluctuations, that is, entropic forces (in the absence of direct attractive protein-protein interactions). However, if the curvature induced by the pro- teins is large enough, bending contributions to the free energy of the system can lead to an 3 effective attraction between proteins and to the formation of spatially inhomogeneous pat- terns in protein distribution and membrane curvature. The details of membrane-mediated protein-protein interactions have been thoroughly studied in the past. [15 -- 18] Furthermore, we will consider the possibility of geometric constraints on the membrane, such as the tether- ing of the membrane to a rigid cell wall/cortex or the existence of a membrane area reservoir at non-zero tension. Interestingly, it was recently shown that solid particles such as proteins can sense the local membrane curvature imposed by geometric constraints on the membrane. [19] FIG. 1: The proteins (yellow) impose a spontaneous curvature Cp on the membrane (blue). De- pending on the interplay between curvature, entropic, and membrane tethering/tension forces, proteins might repel, resulting in a spatially homogeneous spherical membrane (turquoise), or they might attract, leading to the spontaneous formation of inhomogeneous patterns of membrane curvature and protein density. Here, we have found that, in realistic situations, spontaneous pattern formation can be induced either by an increase in the surface density of curvature-inducing proteins, or by a decrease in the strength of the geometric constraints on the membrane. Furthermore, these two parameters can also control the size and number of protein-rich (highly curved) and protein-poor (weakly curved) domains. These mechanisms could be exploited by cells in order to trigger spatial organisation of the plasma membrane on demand, and could in principle be replicated in artificial model systems. The paper is organised as follows. In section II, we present the continuum model for the energetics and dynamics of the system, and examine the linear stability of the dynamical 4 repulsionattractionCp equations for the shape of the membrane and the protein density distribution. In section III, we explore spontaneous pattern formation in the system as a function of all relevant param- eters. Finally, in section IV we discuss the applicability and consequences of our results in real biological or biomimetic systems. II. METHODS A. Energetics We will adopt a continuum elastic model of a closed membrane, which might represent a model vesicle or a biological cell, and study the stability of spherical shapes to perturbations in the presence of curvature-inducing proteins that decorate the membrane. The shape of a quasi-spherical membrane can be written in spherical coordinates as R(θ, φ) = R[1 + u(θ, φ)]r, where R is the radius of the unperturbed sphere, u(θ, φ) is a scalar function that describes the deviations from the sphere, and r is the radial unit vector, see figure 2. The distribution of proteins on the membrane can be described in a similar way, with the surface number density ρ(θ, φ) = ρ0[1 + ψ(θ, φ)]. Here, ρ0 is the average protein number density, i.e. ρ0 = N/4πR2 if N is the total number of proteins on the membrane, and the function ψ(θ, φ) represents the deviations from a homogeneous distribution of proteins. We will assume that each protein covers a patch of membrane of area a0, and imposes a spontaneous curvature Cp on the membrane, see figure 1. The bending free energy of the (cid:90) Fb = κ 2 membrane can then be written within the spontaneous curvature model [20 -- 23] as dA [C 2 − 2Cpρa0C] (1) where κ is the bending rigidity of the membrane, and C is the local membrane curvature, with C = C1 + C2, where C1 and C2 are the two principal curvatures. The second term inside the integral represents the simplest possible coupling between protein density and local curvature. It can also be interpreted as a position-dependent spontaneous curvature C0(θ, φ) ≡ Cpρ(θ, φ)a0, which varies from C0 = 0 in the absence of proteins, with ρ = 0, to C0 = Cp for full coverage of proteins, with ρ = 1/a0. The local membrane curvature C(θ, φ) can be written explicitly as a function of u(θ, φ), as described in ref 24. Besides the bending contributions to the free energy, we need to take into account the entropic contributions due to the mixing and density fluctuations of the proteins. To lowest 5 FIG. 2: The shape of the almost spherical membrane is described by a vector function R(θ, φ), whereas the protein distribution is described by a scalar function ρ(θ, φ) represented by the colour- coding, e.g. yellow and blue could correspond to high and low protein density, respectively. order, this contribution to the free energy can be incorporated as dA [ξ2(∇ψ)2 + ψ2] (2) (cid:90) Fd = 1 2χ Here, χ and ξ are the compressibility and the correlation length of the protein density fluctuations, respectively. The first term in the integral penalises the creation of interfaces between high protein density and low protein density regions, whereas the second term penalises deviations from a homogeneous protein distribution. We will also consider the effect of the tethering of the membrane to a cell wall or acto- myosin cortex, by including a harmonic confinement potential of the form (cid:90) Fh = kteR2 2 dA u2 (3) where kte is an effective spring constant per unit area, which in general may include con- tributions from specific interactions (i.e. proteins that directly link the membrane to the wall/cortex) as well as non-specific interactions such as steric repulsion, van der Waals attraction or electrostatic attraction/repulsion. Within this effective description, the cell wall/cortex is taken to be spherical and rigid (i.e. much more rigid than the membrane), and kte penalises deviations of the membrane position from the (optimal) equilibrium membrane- wall distance. 6 Lastly, we consider the possibility that the membrane is connected to a membrane area reservoir at constant membrane tension. A constant membrane tension is typical of biological cells, [25, 26] and can be mimicked in model vesicle systems by the use of micropipette aspiration. The contribution of a membrane tension σ to the free energy is Ft = σ dA (4) The total free energy can finally be written as the sum of these four contributions, with (cid:90) dA F F = Fb + Fd + Fh + Ft = with the free energy density F ≡ κ 2 [C 2 − 2Cpρa0C] + 1 2χ [ξ2(∇ψ)2 + ψ2] + kR2 2 u2 + σ In addition, we will explicitly impose constraints on the volume enclosed by the membrane (cid:90) (cid:90) (5) (6) (7) (8) (representing osmotic balance), so that 4πR3 3 = as well as on the total number of proteins N on the membrane, so that dV (cid:90) dA ρ N = ρ04πR2 = at all times. B. Dynamics The effective force exerted on the membrane in the radial direction will be balanced by a frictional force, leading to a dynamical equation for the shape of the membrane as a function of time t ∂tu(θ, φ, t) = −Lu δF δu(θ, φ) (9) where Lu is a transport coefficient corresponding to the membrane mobility. On the other hand, the dynamical equation describing the diffusion of the proteins on the membrane can be written in the form of a continuity equation ∂tψ(θ, φ, t) + ∇ · J = 0 7 (10) with a current density J = −Lψ∇µ, where Lψ is another transport coefficient and µ = δF/δψ(θ, φ) is the chemical potential. Putting all together, the dynamical equation for the protein density becomes ∂tψ(θ, φ, t) = Lψ∇2 (cid:18) δF (cid:19) δψ(θ, φ) (11) The Laplacian operator on a sphere can be written as ∇2 ≡ − 1 R2 L2, with the operator − L2 ≡ 1 sin θ ∂θ(sin θ∂θ) + 1 sin2 θ ∂2 φ (12) This operator is diagonal in the basis of spherical harmonics Y(cid:96)m(θ, φ). In particular, it satisfies L2Y(cid:96)m(θ, φ) = (cid:96)((cid:96) + 1)Y(cid:96)m(θ, φ) (13) C. Linear stability analysis To leading order in u and ψ, and taking into account the constraints (7 -- 8) on the enclosed volume and total number of proteins on the membrane, we can write equations (9) and (11) as and ∂tu(θ, φ, t) = −Lu (cid:20)(cid:16) κ R2 (cid:17)(cid:16) L2 − 2 (cid:17) (cid:18) ξ2 (cid:19) L2 + σ (cid:20) 1 χ ∂tψ(θ, φ, t) = −Lψ R2 L2ψ + 1 χ L4ψ + κCpa0ρ0 R R2 u + kteR2u + κCpa0ρ0 (cid:21) (cid:16) L2 − 2 (cid:17) (cid:21) L2(cid:16) L2 − 2 (cid:17) u ψ R (14) (15) We can write the solutions u(θ, φ, t) and ψ(θ, φ, t) as a sum of spherical harmonics, which provide a complete set of orthogonal functions on the sphere, so that u(θ, φ, t) = u(cid:96)m(t)Y(cid:96)m(θ, φ) and ψ(θ, φ, t) = ψ(cid:96)m(t)Y(cid:96)m(θ, φ) (16) (cid:96),m (cid:96),m where u(cid:96)m and ψ(cid:96)m are the amplitudes of the corresponding modes, and we have (cid:96) = 0, 1, 2... and m ≤ (cid:96). However, the constraints (7 -- 8) on the enclosed volume and total number of proteins on the membrane imply that the zero-amplitudes u00 and ψ00 cannot be varied independently. Explicitly imposing these constraints results in expressions for u00 and ψ00 as a function of the squared amplitudes of all modes (cid:88) (cid:88) 4πu00 = −(cid:88) √ (cid:96),m 8 u2 (cid:96)m (17) √ 4πψ00 = (cid:88) (cid:96),m 1 2 [2 − (cid:96)((cid:96) + 1)]u2 (cid:96)m − 2 (cid:88) (cid:96),m u(cid:96)mψ(cid:96)m (18) Equations (17) and (18) imply that u00 and ψ00 are a function of the higher-order amplitudes, and furthermore, that they are of quadratic order (they are equal to a sum of u2 (cid:96)m and u(cid:96)mψ(cid:96)m terms). For this reason, the u00 and ψ00 terms are negligible to linear order, and we can rewrite (16) as u(θ, φ, t) (cid:39) (cid:88) (cid:96)≥1,m u(cid:96)m(t)Y(cid:96)m(θ, φ) and ψ(θ, φ, t) (cid:39) (cid:88) (cid:96)≥1,m ψ(cid:96)m(t)Y(cid:96)m(θ, φ) (19) Inserting (19) into (14) and (15), we can rewrite the dynamical equations as separate equations for each of the (cid:96) ≥ 1 modes. Introducing a rescaled time variable τ ≡ as well as dimensionless parameters (cid:18) κLu (cid:19) R2 t (20) K ≡ kteR4 κ , T ≡ σR2 κ the equations become , P ≡ ξ2 R2 , M ≡ Lψ κLuχ , S ≡ ρ0a0CpR, and B ≡ κχ R2 (21) − ∂τ u(cid:96)m = {[(cid:96)((cid:96) + 1) + T ] ((cid:96) + 2)((cid:96) − 1) + K} u(cid:96)m + S((cid:96) + 2)((cid:96) − 1)ψ(cid:96)m (22) and − (1/M )∂τ ψ(cid:96)m = BS(cid:96)((cid:96) + 1)((cid:96) + 2)((cid:96) − 1)u(cid:96)m + (cid:96)((cid:96) + 1) [1 + P (cid:96)((cid:96) + 1)] ψ(cid:96)m (23) The solutions to (22) and (23) will have the form u(cid:96)m(τ ) = u(cid:96)m(0)eλτ , ψ(cid:96)m(τ ) = ψ(cid:96)m(0)eλτ (24) Inserting these solutions back into (22) and (23), and setting the determinant of the coeffi- cients to zero, we can obtain an equation for the growth rates λ of the characteristic modes of the system, which reads with coefficients λ2 + bλ + c = 0 b ≡ K − 2T + (cid:96)((cid:96) + 1) [M + T − 2 + (cid:96)((cid:96) + 1)(M P + 1)] 9 (25) (26) and c ≡ M (cid:96)((cid:96) + 1)(cid:8)[K + ((cid:96)((cid:96) + 1) + T ) ((cid:96) + 2)((cid:96) − 1)] [1 + P (cid:96)((cid:96) + 1)] − W ((cid:96) + 2)2((cid:96) − 1)2(cid:9) where we have defined the parameter W ≡ BS2 = κχρ2 0a2 0C 2 p (27) (28) The two characteristic modes of the system given by the solutions to (25) can finally be (cid:104) − K + 2T − (cid:96)((cid:96) + 1) [M + T − 2 + (cid:96)((cid:96) + 1)(M P + 1)] (cid:105) (cid:114)(cid:104) (cid:105)2 K − 2T − (cid:96)((cid:96) + 1) [M − T + 2 + (cid:96)((cid:96) + 1)(M P − 1)] written as 1 λ± = 2 ± 1 2 + 4M W (cid:96)((cid:96) + 1)((cid:96) + 2)2((cid:96) − 1)2 (29) Because b in (25) always satisfies b > 0 for all modes with (cid:96) ≥ 1, we know that the amplitude with the smaller value, λ−, is always negative for all (cid:96)-modes. On the other hand, the larger one, λ+, might be positive or negative depending on the (cid:96)-mode and on the values of the parameters W , K, T , P , and M . It is also worth noting that the stability analysis is independent of the value of m of the spherical harmonics. This ultimately arises from the fact that the eigenvalues of the Laplacian of a spherical harmonic are independent of its m-value. The physical significance of the five dimensionless parameters is the following. The pa- rameter W represents the protein-induced spontaneous curvature, and increases both with the average density ρ0 of proteins on the membrane and with the characteristic spontaneous curvature Cp of these proteins. The parameter K represents the strength of the confine- ment of the membrane by its interaction with the rigid cell wall/cortex. The parameter T represents the magnitude of the membrane tension. The parameter P compares the corre- lation length of the protein density fluctuations to the size of the cell or vesicle. Given that correlation lengths are typically of the order of nanometers whereas cell or vesicle sizes are of the order of micrometers, P will generally be small, and will decrease or increase with increasing or decreasing cell/vesicle size, respectively. Finally, the parameter M compares the typical timescale of the changes in membrane shape (R2/Luκ) to that of changes in protein distribution (χR2/Lψ). Importantly, we note that all five dimensionless parameters are always positive. 10 III. RESULTS A positive value of the mode amplitude λ+ implies that fluctuations of this mode will grow instead of decaying, and therefore modes with λ+ > 0 are unstable. If, by small changes in one of the system parameters W , K, T , P , or M , one of the modes λ+ switches from having a negative value to having a positive value, the system will exhibit spontaneous pattern formation. In the following, we will explore the conditions under which spontaneous pattern formation occurs. First of all, we note that, as described above, the (cid:96) = 0 mode cannot vary independently as it is fixed by the constraints on the enclosed volume and total number of proteins, see (17 -- 18). Furthermore, by substituting (cid:96) = 1 in (29), we find the mode amplitudes −K and −2M (1 + 2P ), which can never be positive, implying that the (cid:96) = 1 mode can never become unstable. It can, however, become marginally stable in the particular case of K = 0, i.e. in the absence of tethering to the cell wall. This reflects the fact that (cid:96) = 1 deformations of the membrane shape are equivalent to spatial translations, and that the curvature energy of the membrane is invariant to such translations. The presence of the cell wall, however, breaks translational invariance. All things considered, instabilities and therefore spontaneous pattern formation can occur only for higher modes (cid:96) ≥ 2, which we will discuss below. The larger solution λ+ of (25) will be positive, with λ+ > 0, if and only if c < 0. Using the definition of c in (27), this condition can be rewritten as {K + [(cid:96)((cid:96) + 1) + T ]((cid:96) + 2)((cid:96) − 1)}[1 + P (cid:96)((cid:96) + 1)] W > ((cid:96) + 2)2((cid:96) − 1)2 ≡ W(cid:96) (30) which serves as a definition of W(cid:96), the critical value of the parameter W above which mode (cid:96) becomes unstable. Going back to the definition of W in (28), the inequality (30) implies that an increase in the average density ρ0 of curvature-inducing proteins beyond a critical density will trigger an instability with spherical harmonic mode (cid:96) in both the shape and protein distribution of the membrane. Furthermore, the critical protein density that is needed to trigger an instability decreases with increasing protein spontaneous curvature Cp. Importantly, we note that the critical value W(cid:96) is independent of the parameter M , and therefore depends only on three parameters, P , K, and T . In fact, the parameter M drops out of all relevant equations in the following, so that pattern formation in the system turns out to be governed by only four dimensionless parameters: W , K, T , and P . This is a consequence of the fact that M is a mobility parameter that related the timescale of 11 changes in membrane shape to that of changes in protein distribution, and as such it only affects the dynamics of the system. Alternatively, the instability condition c < 0 can be written as K < W ((cid:96) + 2)2((cid:96) − 1)2 1 + P (cid:96)((cid:96) + 1) − [(cid:96)((cid:96) + 1) + T ]((cid:96) + 2)((cid:96) − 1) ≡ K(cid:96) or W ((cid:96) + 2)((cid:96) − 1) 1 + P (cid:96)((cid:96) + 1) − T < K ((cid:96) + 2)((cid:96) − 1) − (cid:96)((cid:96) + 1) ≡ T(cid:96) (31) (32) which define K(cid:96) and T(cid:96), the critical values of K and T , respectively, below which mode (cid:96) becomes unstable. Going back to the definitions of K and T in (21), the inequalities (31) and (32) respectively imply that the shape and protein distribution instability can also be triggered by a decrease in the tethering strength of the membrane to the cell wall/cortex, or by a decrease in the membrane tension. Once again, we note that the critical values K(cid:96) and T(cid:96) are independent of the parameter M . As outlined in the previous two paragraphs, the parameters that could presumably be actively controlled by a biological cell or tuned in experiments with model vesicles are W , i.e. the density of proteins on the cell surface, K, i.e. the tethering strength of the membrane to the cell wall/cortex, and T , the membrane tension. The parameter P , on the other hand, represents the correlation length of the protein density fluctuations, i.e. the typical distance at which proteins can sense each other, and will in general be fixed for a given system. It therefore makes sense to explore the behaviour of the system when W , K, and T are varied for a fixed value of P . Using (30), in figure 3 we have plotted the lines W = W(cid:96)(K) for (cid:96) ≥ 2, using T = 0 (i.e. negligible membrane tension) and three different values of P , namely P = 0.1, 0.02, and 0.005. For a vesicle/cell of radius 1 µm, these values of P would correspond to correlation lengths of ξ = 320 nm, 140 nm, and 70 nm, respectively. In the region of low W and high K, depicted in grey, the spherical state with a homogeneous protein distribution is stable. As W is increased from low values, the system will hit the instability of the first unstable mode, with a given value of (cid:96) which will depend on the value of K. Alternatively, if K is decreased from high values, the system will also hit the instability of the first unstable mode with a given (cid:96) which will depend on the value of W . The higher the value of K, the higher the value of (cid:96) of the first unstable mode as W is increased. Similarly, the higher the value of W , the higher the value of (cid:96) of the first unstable mode as K is decreased. 12 There are important differences in the way in which W and K act to trigger pattern formation. Independently of the value of K, and even for K = 0, a sufficiently high W will always lead to pattern formation. On the other hand, a decrease in K can only lead to pattern formation if W is above the critical value W(cid:96)(K = 0). Furthermore, we note that figure 3 has a semilogarithmic axis: whereas the critical value of W above which pattern formation occurs is always in the vicinity of 1, with W (cid:38) 1, the critical value of K below which pattern formation occurs can vary over many orders of magnitude. Pattern formation is therefore particularly sensitive to W , i.e. to the density of curvature-inducing proteins on the membrane. And what is the effect of P , that is, of the correlation length of the protein density fluctuations? Let us now compare figures 3(a), (b), and (c). For the highest value of P , in (a), the first unstable mode for increasing W at vanishing K is (cid:96) = 2, whereas larger values of K lead to the instabilities of higher-order modes with (cid:96) > 2. As P is decreased, as in (b), the first unstable mode at vanishing K is now (cid:96) = 3: the mode (cid:96) = 2 is not the first unstable mode for any value of K. When P is decreased even further, as in (c), (cid:96) = 4 becomes the first unstable mode at vanishing K, and neither (cid:96) = 2 nor (cid:96) = 3 are the first unstable modes for any value of K. This trend continues as P is decreased further, with progressively higher order modes becoming the first unstable mode at vanishing K. Moreover, we note that, as P is decreased, the critical value of W above which pattern formation occurs moves closer and closer to W = 1. In figure 3 we have explored the stability behaviour of the system as a function of W and K, for fixed vanishing tension T = 0. Considering a fixed non-zero tension T > 0 leads to the same qualitative behaviour of the system as a function of W and K. Furthermore, the behaviour of the system as a function of W and T for fixed K is qualitatively identical to that as a function of W and K for fixed T , leading to instability lines analogous to those in figure 3. We thus omit these results for the sake of brevity. As just described, in order to characterise the system, it is particularly important to identify the first unstable mode when W is increased, that is, the mode with smallest W(cid:96) for given values of P , K, and T , which we will denote as (cid:96)∗ W . The critical value of W ≡ min(cid:96)(W(cid:96)). The above which the first unstable mode becomes unstable is then W ∗ ≡ W(cid:96)∗ boundaries between the regions in the three-dimensional (P, K, T ) parameter space in which W modes (cid:96) and (cid:96) + 1 are the first unstable mode for increasing W can be obtained from the 13 FIG. 3: Instability lines W = W(cid:96)(K) for modes (cid:96) ≥ 2, vanishing tension T = 0, and three values low number of curvature- of P : (a) P = 0.1, (b) P = 0.02, and (c) P = 0.005. For low W (i.e. inducing proteins) and high K (i.e. strong confinement of the membrane due to tethering to the cell wall/cortex), the spherical homogeneous state is stable (grey region). As W is increased or K is decreased, the system will hit an instability with a given value of (cid:96). If W or K keep increasing or decreasing, respectively, they will hit the instabilities of further modes. The higher the value of K or W , the higher the value of (cid:96) of the first unstable mode as W is increased or K is decreased. The parameter P represents the correlation length of the protein density fluctuations. condition W(cid:96) = W(cid:96)+1, which can be written explicitly using (30) as P = ((cid:96) − 1)(cid:96)((cid:96) + 2)((cid:96) + 3)(2 + T ) − 2K[1 − (cid:96)((cid:96) + 2)] ((cid:96) − 1)(cid:96)((cid:96) + 2)((cid:96) + 3){[(cid:96)((cid:96) + 2) − 4]((cid:96) + 1)2 − 2T} − K[(cid:96)((cid:96) + 1)2((cid:96) + 2) − 4] (33) In figure 4, we have used equation (33) to explore pattern formation in (a) the (P, K, T = 0) plane and (b) the (P, K = 0, T ) plane. For any point in (P, K, T ) space, we can obtain the critical value W ∗ above which pattern formation occurs, using (30). This information is also colour-coded in figure 4. Several important observations can be made: (i) Once again, we see that K and T have qualitatively similar effects in pattern formation. (ii) Both an increase in K or T , as well as a decrease in P lead to increasingly higher-order modes being the first unstable mode. (iii) In most regions of the parameter space, the critical value W ∗ above which pattern formation occurs is very close to 1. The only exception is the region of P ≈ 1 and large K or T , in which W ∗ can be much larger than one. A particularly important case, with regards to its experimental relevance, is that of a model lipid vesicle, for which we have both K = 0 (there is no wall or cortex attached to the membrane) and T = 0 (if we are considering a flaccid, unstretched vesicle). This corresponds to the bottom part of of both figure 4(a) and (b). In this limit case, which mode first becomes unstable when W (i.e. the number of curvature-inducing proteins on 14 991011(a)(b)(c) the membrane) is increased depends only on the parameter P (i.e. the correlation length of the protein density fluctuations), with the boundaries between (cid:96) and (cid:96) + 1 being the first unstable modes given by the simple expression P = 2 [(cid:96)((cid:96) + 2) − 4]((cid:96) + 1)2 (34) as obtained from equation (33) with K = T = 0. Using equation (34), we predict that for a tensionless spherical vesicle, the (cid:96) = 2 mode will be the first unstable mode if P > 1/18, the (cid:96) = 3 mode will be the first unstable mode if 1/18 > P > 1/88, the (cid:96) = 4 mode if 1/88 > P > 1/250, and so on. For a typical vesicle of radius 1 µm and a typical correlation length of ξ = 20 nm, we have P = 4· 10−4, and we find that the (cid:96) = 8 mode will be the first unstable mode. FIG. 4: Pattern formation triggered by an increase in the number of curvature-inducing proteins: First unstable modes (cid:96)∗ W when W is increased, (a) as a function of the parameters P and K for T = 0; and (b) as a function of the parameters P and T for K = 0. The critical value W ∗ above which pattern formation occurs can be calculated from (30), and is indicated by the colour-coding, which is the same for (a) and (b). The vertical dashed lines in (a) correspond to the three particular cases P = 0.1, P = 0.02 and P = 0.005 displayed in figure 3. Alternatively, we could ask ourselves what is the first unstable mode (cid:96)∗ K when K is decreased for given values of P , W , and T , or equivalently, the mode with largest K(cid:96) for given P , W , and T . The critical value of K below which the first unstable mode becomes 15 (b)(a) unstable is then K∗ ≡ K(cid:96)∗ ≡ max(cid:96)(K(cid:96)). The boundaries between the regions in the three- dimensional (P, W, T ) parameter space in which modes (cid:96) and (cid:96) + 1 are the first unstable K mode for decreasing K can be obtained from the condition K(cid:96) = K(cid:96)+1, which can be written explicitly using (31). The resulting (P, W ) stability diagram for the particular case In the same way, we can find the first unstable mode of T = 0 is shown in figure 5(a). (cid:96)∗ T when T is decreased for given values of P , W , and K, with a critical value given by ≡ max(cid:96)(T(cid:96)), and boundaries in the (P, W, K) parameter space given by T(cid:96) = T(cid:96)+1, T ∗ ≡ T(cid:96)∗ which can be written explicitly using (32). The resulting (P, W ) stability diagram for the T particular case of K = 0 is shown in figure 5(b). Once again, we find that K (the strength of the tethering of the membrane to the cell wall/cortex) and T (the membrane tension) behave in a qualitatively similar way. As ex- pected from figure 3, an instability can only occur for decreasing K (or T ) if W is sufficiently high. This minimum value of W required for pattern formation approaches W = 1 for small P . Indeed, figure 5 illustrates very clearly a striking feature of the system: for low values of P (which are the most typical given that the correlation length of protein density fluctu- ations ξ is normally much smaller than the membrane radius R), values of W only slightly above 1 can lead to the instability of modes with very high (cid:96) when K or T are decreased. This is evidenced by the high density of boundary lines in the region of low P (cid:28) 1, W (cid:38) 1. IV. DISCUSSION A. Estimation and control of model parameters in real systems We have shown above that pattern formation in a spherical membrane containing curvature-inducing proteins is controlled by the four dimensionless parameters W , K, T and P , which represent the number of curvature-inducing proteins on the membrane, the strength of the membrane tethering to the cell wall/cortex, the membrane tension, and the correlation length of protein fluctuations, respectively. An important question is then: what are the typical values of these parameters in real systems, and to what extent can they be controlled by a biological cell, or tuned in experiments with model vesicles? The parameter to which the system is most sensitive is W , see figures 3, 4 and 5. Even if K, T , and P vary across many orders of magnitude, the critical value W above which pattern 16 (a) Pattern formation triggered by a decrease in the tethering strength of the membrane to FIG. 5: the cell wall/cortex: First unstable modes (cid:96)∗ K when K is decreased, as a function of the parameters P and W for T = 0. The critical value K∗ below which pattern formation occurs can be calculated from (31), and is indicated by the colour-code. The vertical dashed lines correspond to the three particular cases P = 0.1, P = 0.02 and P = 0.005 displayed in figure 3. (b) Pattern formation triggered by a decrease in membrane tension: First unstable modes (cid:96)∗ T when T is decreased, as a function of the parameters P and W for K = 0. The critical value T ∗ below which pattern formation occurs can be calculated from (32), and is indicated by the colour-code. formation occurs always stays in the proximity of W ∗ ≈ 1, except in the extreme case of very high K (or T ) and P ≈ 1 simultaneously, see figure 4. Going back to the definition of W in terms of the dimensionful system parameters in (28), we see that the requirement W ≥ W ∗ ≈ 1 implies that κχρ2 (cid:38) 1. Here, κ is the bending rigidity of the membrane, χ is the compressibility of the protein density fluctuations, ρ0 is the average density of 0C 2 p 0a2 curvature-inducing proteins on the membrane, a0 is the lateral area of a single protein, and Cp is the protein spontaneous curvature. Furthermore, in the limit of low protein density ρ0a0 (cid:28) 1, the compressibility χ can be approximated by χ ≈ 1/(kBT ρ0), [27] where kB is Boltzmann's constant and T is the temperature. In this limit, the requirement for pattern (cid:38) 1. In general, the protein spontaneous curvature √ a0, so that we pa0 ≈ 1. We finally conclude that pattern formation typically occurs for average Cp will be of the order of the (inverse) characteristic length of the protein formation thus becomes (κ/kBT )ρ0a2 0C 2 p can take C 2 protein densities satisfying ρ0a0 (cid:38) kBT /κ 17 (35) no instabilityno instability(b)(a) Here, ρ0a0 is simply the dimensionless area fraction of membrane covered by the protein. Typical values of the bending rigidity of membranes range from 10 to 100 kBT , leading to a critical protein coverage of the order of ρ0a0 ∼ 0.01 -- 0.1. Importantly, the range obtained self- consistently validates the low protein density assumption made above. Furthermore, such coverages are within the range achievable both in biological cells as well as in model vesicles. In this picture, a biological cell could up- or down-regulate the expression of the curvature- inducing protein in order to switch between patterned and non-patterned conformations, see figure 4. Furthermore, the concentration of curvature-inducing proteins on the membrane could be directly controlled in experiments with model vesicles. Let us now turn to the dimensionless parameters K ≡ kteR4/κ and T ≡ σR2/κ, which both act as geometric constraints on the membrane: K represents the confinement of the membrane due to its interaction/tethering to the cell wall or cortex, whereas T represents the membrane tension, which acts to minimise the cell membrane area. It is interesting to note that, while model membranes such as Giant Unilamellar Vesicles show clear shape fluctuations due to thermal excitation of bending modes, [28] eukaryotic cells or bacteria do not show such fluctuations. The latter is an indication that, in such systems, membrane confinement and tension must overpower bending, and consequently that in these systems K (cid:29) 1 and/or T (cid:29) 1, as can be confirmed by the quantitative estimates that follow. Estimates of the confinement strength kte of biological membranes due to the interaction with the corresponding cell wall or cortex do not abound in the literature. In Ref. 29, the density of membrane-cortex linkers in eukaryotic cells was estimated to be around ρlink ≈ 100 µm−2, whereas the spring constant of a typical linker was estimated to be klink ≈ 10−4 N/m. The effective tethering strength should go as kte ≈ ρlinkklink, leading to the estimate kte ≈ 1010 J/m4 ≈ 2.5 · 106 kBT /µm4. Considering a typical range of bending rigidities κ = 10 -- 100 kBT , and a typical cell radius ranging from from R = 1 µm to 10 µm, we find values for K ranging from 2.5 · 104 up to 2.5 · 109. Cells could then actively switch between patterned and non-patterned conformations by down- or up-regulating the concentration of linker proteins between the plasma membrane and the cell wall/cortex, see figure 5(a). The typical tension of cellular membranes, on the other hand, has been extensively mea- sured for different cell types, and can range from σ = 3 pN/µm for epithelial cells up to about σ = 300 pN/µm for keratocytes. [25, 26] Using the range σ = 3 -- 300 pN/µm for the 18 membrane tension, together with the estimates κ = 10 -- 100 kBT for the bending rigidity of the membrane and R = 1 -- 10 µm for a typical cell radius, we obtain values of T ranging from 7 to 7 · 105. Cells can actively regulate their own tension in order to maintain homeosta- sis. [25] In this way, cells could switch between patterned and non-patterned conformations by actively decreasing or increasing the tension of their plasma membrane, see figure 5(b). Furthermore, triggering of pattern formation via a decrease in membrane tension could be explored in experiments using model Giant Unilamellar Vesicles aspirated by micropipettes, which allows direct experimental control over the membrane tension. Let us finally examine the dimensionless parameter P , defined as P ≡ ξ2/R2, where ξ is the correlation length of the protein density fluctuations and R is the radius of the cell or vesicle. The correlation length ξ is a measure of the distance at which proteins or protein clusters can sense each other, typically via membrane-mediated interactions in the absence of other long-ranged interactions. Previous work [15 -- 18] has shown that the typical length scale of membrane-mediated interactions is the size of the curvature-inducing element itself, so that we can use an estimate of ξ = 10 -- 20 nm. On the other hand, the radius of cells or cellular compartments, as well as of model vesicles, can range between R = 100 nm and 10 µm. With this, we find a range of P ∼ 10−6 -- 10−2. In this range of values with P (cid:28) 1, as described above, pattern formation is tighly controlled by the number of curvature-inducing proteins, with an instability occurring as soon as W (cid:38) 1. Moreover, the value of P directly controls the (cid:96)-order of the first unstable during pattern formation, and as a consequence controls the typical size of the protein-rich, highly-curved domains. The consequences of this fact are discussed in the following section. B. Biological relevance 1. Cell division Cell division requires polarisation of the cell, so that the spherical symmetry of the cell is broken, leading to two identifiable poles as well as an equatorial line. Spontaneous pattern formation via an instability due to the presence of curvature-inducing proteins, as described here, provides a simple mechanism for such a symmetry breaking. This occurs for the mode (cid:96) = 2 of the instability, see figure 6(a), which can be the first unstable mode as long as 19 P > 1/18 (cid:39) 0.056, as determined from (34) and displayed on figures 4 and 5. For a typical cell size of R = 1 µm, such values of P would correspond to a correlation length ξ for protein density fluctuations larger than 230 nm. This value appears too high for a typical protein, given that the correlation length is expected to be of the order of the protein size, i.e. a couple of tens of nanometers. It could, however, be a plausible value for protein clusters, composed of a few tens of proteins with lateral sizes of the order of 100 nm. If such clusters arose by a separate mechanism, a curvature-instability such as the one described here could lead to cell polarisation. Several proteins, many of them related to cell division, are known to preferentially localise at the poles of bacterial membranes. [30 -- 34] We note, however, that the generic mechanism proposed here is distinct and unrelated to the well-studied Min system, which serves to localise the FtsZ protein ring in rod-shaped bacteria. [10, 11] The Min system involves both membrane bound as well as cytosolic components, and locates the bacterial equator via an oscillatory mechanism. The mechanism proposed here might however explain why FtsZ proteins can spontaneously self-assemble on vesicles, even in the absence of the Min system. [8, 9] In eukaryotic cells, cell division is mediated by the cytoskeleton, in particular by the mitotic spindle and the cleavage furrow. The mechanism underlying the initial positioning of this cell division apparatus is however not well understood at the molecular level. [35, 36] Even if the mechanism described here did not play a direct role in cell division, it pro- vides a generic pathway for symmetry breaking and the initiation of division of spherical membranes into two equally sized daughters, using only a minimal number of ingredients. As such, it could serve as a plausible mechanism for the division of protocells, as well as of synthetic cells in bottom-up synthetic biology. [37 -- 39] 2. Large-scale protein organisation Going beyond (cid:96) = 2, the mechanism described here also predicts pattern formation with modes of intermediate (cid:96), e.g. in the range (cid:96) = 5 -- 10. Formation of such patterns would imply protein-rich, strongly-curved clusters with sizes on the order of 1/5 to 1/10 of the cell size, see figure 6(b). Such patterns have been observed in L-form [6, 7] bacteria, as well as in coccal bacteria. [5] The patterns formed by PlsY and CdsA proteins (both essential to lipid metabolism) in Staphylococcus aureus, in particular, show a striking coupling between 20 protein density and membrane curvature. [12] As seen in figures 4 and 5, and determined from (34), modes with (cid:96) ≤ 10 are expected for P > 1.4· 10−4 which, for a typical cell size of R = 1 µm, would correspond to correlation lengths ξ > 10 nm, well within the biologically plausible range. 3. Nano-sized membrane rafts The existence of protein-rich raft domains in the plasma membrane was controversial for some time, partly due to a conflation between the macroscale fluid-fluid phase separation observed in model lipid membranes with the observation of rafts in living cells. [40] Nev- ertheless, it is currently accepted that rafts are dynamic, fluctuating assemblies of proteins with sizes on the order of tens of nanometers. [40 -- 43] The precise physical mechanism behind raft formation, however, is still a matter of debate. Currently proposed theories include that rafts are compositional fluctuations near the critical point of fluid-fluid phase separation in lipid membranes, [44] or that the actin cortex underlying the plasma membrane acts as a 'picket-fence' which inhibits the lateral diffusion of proteins and promotes the formation of nano-scale aggregates. [45] The model that we have presented here predicts that, under biologically reasonable pa- rameters, curvature-inducing proteins can spontaneously self-organise into patterns that may be built from spherical harmonics with very high-order (cid:96)-modes, with (cid:96) (cid:29) 1. As a conse- quence, in such cases the typical size of the protein-rich domains (which goes as ∼ R/(cid:96)) will be much smaller than the cell size R, leading to domain sizes on the order of tens of nanometers for a micron-sized cell, see figure 6(c) for an example with (cid:96) = 100. It is therefore tempting to speculate that the mechanism presented here might also be connected to the existence of such nano-scale protein-rich rafts. Indeed, let us use the quantitative estimates of parameters obtained above, with typical values of membrane bending rigidity κ = 10 kBT , correlation length of protein density fluctuations of ξ = 10 nm, tethering strength of the membrane to the cell cortex kte ≈ 2.5· 106 kBT /µm4, and membrane tension σ = 30 pN/µm. For a small cell of radius 1 µm, we can calculate our dimensionless parameters as P = 10−4, K = 2.5 · 105, and T = 7 · 102. Using these values in equation (33), we expect the first unstable mode to be (cid:96) ≈ 50. This would correspond to a typical domain size R/(cid:96) ≈ 20 nm. For a larger cell of radius 10 µm, we calculate P = 10−6, K = 2.5·109, and T = 7·104. Using 21 these values in (33), we find (cid:96) ≈ 500 for the first unstable mode, once again corresponding to a typical domain size R/(cid:96) ≈ 20 nm. FIG. 6: Three examples of spontaneous pattern formation via shape and protein distribution instability. (a) Instability (cid:96) = 2, corresponding to cell division with two distinct poles and an equatorial line. The mode (cid:96) = 2 can be triggered if P > 0.056, see figures 4 and 5. (b) Instability (cid:96) = 10, corresponding to large-scale protein organisation such as that observed in Staphylococcus aureus. [12] The mode (cid:96) = 10 can be triggered if P > 1.4 · 10−4. (c) Instability (cid:96) = 100, corresponding to the formation of nano-sized protein-rich membrane rafts. The mode (cid:96) = 100 can be triggered if P > 1.9 · 10−8. C. Summary To summarise, we have explored in detail pattern formation in spherical membranes that contain curvature-inducing proteins. Pattern formation arises from the interplay between membrane curvature energy, protein density fluctuations, and geometric constraints such as membrane tension and confinement forces due to the tethering of the membrane to the cell wall/cortex. We have shown that pattern formation in this system is controlled by just four dimensionless parameters, W , K, T , and P , defined in (21) and (28). These parameters represent the number of curvature-induced proteins on the membrane, the confinement of the membrane due to the cell wall/cortex, the membrane tension, and the correlation length of protein density fluctuations, respectively. In most circumstances, pattern formation is expected to occur as the result of an increase in the average surface density of proteins (i.e. the total number of proteins on the membrane surface), or of a relaxation of the ge- ometric constraints on the membrane due to membrane tension or membrane tethering to the cell wall/cortex. The patterns that arise consist of protein-rich, highly-curved domains 22 10100(a)(b)(c) that alternate with protein-poor, weakly-curved domains. We hypothesise that spontaneous pattern formation as described here might be exploited by biological cells as a way to reg- ulate their geometry in situations that require spatial organisation, symmetry breaking or polarisation of the cell, using only a minimal number of ingredients. Acknowledgements We would like to acknowledge fruitful discussions with J. Garcia-Lara and S. Foster. This work was supported by the Human Frontiers Science Program (HFSP) RGP0061/2013. J.A-C. acknowledges support from the Federal Ministry of Education and Research (BMBF, Germany) via the consortium MaxSynBio, as well as from the Penn State MRSEC, Center for Nanoscale Science, under the award NSF DMR-1420620. Correspondence should be addressed to [email protected]. [1] D. G. Drubin and W. J. Nelson, "Origins of cell polarity," Cell, vol. 84, no. 3, pp. 335 -- 344, 1996. [2] W. J. Nelson, "Adaptation of core mechanisms to generate cell polarity," Nature, vol. 422, no. 6933, pp. 766 -- 774, 2003. [3] A. M. Turing, "The Chemical Basis of Morphogenesis," Philosophical Transactions of the Royal Society of London B: Biological Sciences, vol. 237, no. 641, pp. 37 -- 72, 1952. [4] S. P. Thampi, R. Golestanian, and J. M. Yeomans, "Instabilities and topological defects in active nematics," Europhysics Letters, vol. 105, no. 1, p. 18001, 2014. [5] A. Zapun, T. Vernet, and M. G. Pinho, "The different shapes of cocci," FEMS Microbiology Reviews, vol. 32, no. 2, pp. 345 -- 360, 2008. [6] M. Leaver, P. Dom´ınguez-Cuevas, J. M. Coxhead, R. A. Daniel, and J. Errington, "Life without a wall or division machine in Bacillus subtilis," Nature, vol. 460, no. 7254, pp. 538 -- 538, 2009. [7] R. Mercier, Y. Kawai, and J. Errington, "General principles for the formation and proliferation of a wall-free (L-form) state in bacteria," eLife, vol. 3, no. e04629, 2014. [8] M. Osawa, D. E. Anderson, and H. P. Erickson, "Reconstitution of Contractile FtsZ Rings in 23 Liposomes," Science, vol. 320, no. 5877, pp. 792 -- 794, 2008. [9] R. Shlomovitz and N. S. Gov, "Membrane-mediated interactions drive the condensation and coalescence of FtsZ rings," Physical Biology, vol. 6, no. 4, p. 046017, 2009. [10] J. Schweizer, M. Loose, M. Bonny, K. Kruse, I. Monch, and P. Schwille, "Geometry sensing by self-organized protein patterns," Proceedings of the National Academy of Sciences, vol. 109, no. 38, pp. 15283 -- 15288, 2012. [11] Z. Petr´asek and P. Schwille, "Simple membrane-based model of the Min oscillator," New Journal of Physics, vol. 17, no. 4, p. 043023, 2015. [12] J. Garc´ıa-Lara, F. Weihs, X. Ma, L. Walker, R. R. Chaudhuri, J. Kasturiarachchi, H. Crossley, R. Golestanian, and S. J. Foster, "Supramolecular structure in the membrane of Staphylococ- cus aureus," Proceedings of the National Academy of Sciences of the United States of America, vol. 112, no. 51, pp. 15725 -- 15730, 2015. [13] H. T. McMahon and J. L. Gallop, "Membrane curvature and mechanisms of dynamic cell membrane remodelling," Nature, vol. 438, pp. 590 -- 6, dec 2005. [14] J. Zimmerberg and M. M. Kozlov, "How proteins produce cellular membrane curvature," Nature Reviews Molecular Cell biology, vol. 7, pp. 9 -- 19, jan 2006. [15] R. Golestanian, M. Goulian, and M. Kardar, "Fluctuation-induced interactions between rods on membranes and interfaces," Europhysics Letters (EPL), vol. 33, pp. 241 -- 246, jan 1996. [16] R. Golestanian, M. Goulian, and M. Kardar, "Fluctuation-induced interactions between rods on a membrane," Physical Review E, vol. 54, pp. 6725 -- 6734, dec 1996. [17] T. R. Weikl, M. M. Kozlov, and W. Helfrich, "Interaction of Conical Membrane Inclusions: Effect of Lateral Tension," Physical Review E, vol. 57, no. 6, p. 10, 1998. [18] B. J. Reynwar and M. Deserno, "Membrane-mediated interactions between circular particles in the strongly curved regime," Soft Matter, vol. 7, no. 18, p. 8567, 2011. [19] J. Agudo-Canalejo and R. Lipowsky, "Uniform and Janus-like nanoparticles in contact with vesicles: energy landscapes and curvature-induced forces," Soft Matter, vol. 13, no. 11, pp. 2155 -- 2173, 2017. [20] W. Helfrich, "Elastic properties of lipid bilayers: theory and possible experiments," Zeitschrift fur Naturforschung. Teil C: Biochemie, Biophysik, Biologie, Virologie, vol. 28, no. 11, pp. 693 -- 703, 1973. [21] S. Leibler, "Curvature instability in membranes," Journal de Physique, vol. 47, no. 3, pp. 507 -- 24 516, 1986. [22] U. Seifert, K. Berndl, and R. Lipowsky, "Shape transformations of vesicles: phase diagram for spontaneous-curvature and bilayer-coupling models," Physical Review A, vol. 44, no. 2, pp. 1182 -- 1202, 1991. [23] S. Ramaswamy, J. Toner, and J. Prost, "Nonequilibrium fluctuations, traveling waves, and instabilities in active membranes.," Physical review letters, vol. 84, pp. 3494 -- 3497, 2000. [24] W. Helfrich, "Size distributions of vesicles : the role of the effective rigidity of membranes," Journal de Physique, vol. 47, no. 2, pp. 321 -- 329, 1986. [25] C. E. Morris and U. Homann, "Cell surface area regulation and membrane tension.," The Journal of membrane biology, vol. 179, no. 2, pp. 79 -- 102, 2001. [26] P. Sens and J. Plastino, "Membrane tension and cytoskeleton organization in cell motility," Journal of Physics: Condensed Matter, vol. 27, no. 27, p. 273103, 2015. [27] W. Cai and T. C. Lubensky, "Hydrodynamics and dynamic fluctuations of fluid membranes," Physical Review E, vol. 52, no. 4, pp. 4251 -- 4266, 1995. [28] R. Dimova, S. Aranda, N. Bezlyepkina, V. Nikolov, K. A. Riske, and R. Lipowsky, "A practical guide to giant vesicles. Probing the membrane nanoregime via optical microscopy," Journal of Physics: Condensed Matter, vol. 18, no. 28, pp. S1151 -- S1176, 2006. [29] R. Alert, J. Casademunt, J. Brugu´es, and P. Sens, "Model for Probing Membrane-Cortex Adhesion by Micropipette Aspiration and Fluctuation Spectroscopy," Biophysical Journal, vol. 108, no. 8, pp. 1878 -- 1886, 2015. [30] L. Shapiro, H. H. Mcadams, and R. Losick, "Generating and Exploiting Polarity in Bacteria," Science, vol. 298, no. December, pp. 1942 -- 1946, 2002. [31] E.-M. Lai, U. Nair, N. D. Phadke, and J. R. Maddock, "Proteomic screening and identification of differentially distributed membrane proteins in Escherichia coli.," Molecular microbiology, vol. 52, no. 4, pp. 1029 -- 1044, 2004. [32] S. Thiem, D. Kentner, and V. Sourjik, "Positioning of chemosensory clusters in E. coli and its relation to cell division," The EMBO Journal, vol. 26, no. 6, pp. 1615 -- 1623, 2007. [33] G. R. Bowman, L. R. Comolli, J. Zhu, M. Eckart, M. Koenig, K. H. Downing, W. E. Moerner, T. Earnest, and L. Shapiro, "A Polymeric Protein Anchors the Chromosomal Origin/ParB Complex at a Bacterial Cell Pole," Cell, vol. 134, no. 6, pp. 945 -- 955, 2008. [34] G. Ebersbach, A. Briegel, G. J. Jensen, and C. Jacobs-Wagner, "A Self-Associating Protein 25 Critical for Chromosome Attachment, Division, and Polar Organization in Caulobacter," Cell, vol. 134, no. 6, pp. 956 -- 968, 2008. [35] M. Glotzer, "Cleavage furrow positioning," Journal of Cell Biology, vol. 164, no. 3, pp. 347 -- 351, 2004. [36] F. A. Barr and U. Gruneberg, "Cytokinesis: Placing and Making the Final Cut," Cell, vol. 131, no. 5, pp. 847 -- 860, 2007. [37] M. M. Hanczyc, S. M. Fujikawa, and J. W. Szostak, "Experimental Models of Primitive Cellular Compartments: Encapsulation, Growth, and Division," Science, vol. 302, no. 5645, pp. 618 -- 622, 2003. [38] T. F. Zhu and J. W. Szostak, "Coupled growth and division of model protocell membranes," Journal of the American Chemical Society, vol. 131, no. 15, pp. 5705 -- 5713, 2009. [39] D. Zwicker, R. Seyboldt, C. A. Weber, A. A. Hyman, and F. Julicher, "Growth and division of active droplets provides a model for protocells," Nature Physics, vol. 13, no. 4, pp. 408 -- 413, 2016. [40] K. Jacobson, O. G. Mouritsen, and R. G. Anderson, "Lipid rafts: at a crossroad between cell biology and physics," Nature cell biology, vol. 9, no. 1, pp. 7 -- 14, 2007. [41] D. Lingwood and K. Simons, "Lipid Rafts As a Membrane-Organizing Principle," Science, vol. 327, no. 5961, pp. 46 -- 50, 2010. [42] K. Simons and M. J. Gerl, "Revitalizing membrane rafts: new tools and insights," Nature Reviews Molecular Cell Biology, vol. 11, no. 10, pp. 688 -- 699, 2010. [43] E. Sezgin, I. Levental, S. Mayor, and C. Eggeling, "The mystery of membrane organiza- tion: composition, regulation and roles of lipid rafts," Nature Reviews Molecular Cell Biology, vol. 18, no. 6, pp. 361 -- 374, 2017. [44] S. L. Veatch, P. Cicuta, P. Sengupta, A. Honerkamp-Smith, D. Holowka, and B. Baird, "Crit- ical Fluctuations in Plasma Membrane Vesicles," ACS Chemical Biology, vol. 3, pp. 287 -- 293, may 2008. [45] K. Ritchie, R. Iino, T. Fujiwara, K. Murase, and A. Kusumi, "The fence and picket structure of the plasma membrane of live cells as revealed by single molecule techniques (Review)," Molecular Membrane Biology, vol. 20, no. 1, pp. 13 -- 18, 2003. 26
1010.0727
1
1010
2010-10-04T23:05:48
Intense ultraviolet perturbations on aquatic primary producers
[ "physics.bio-ph", "astro-ph.EP" ]
During the last decade, the hypothesis that one or more biodiversity drops in the Phanerozoic eon, evident in the geological record, might have been caused by the most powerful kind of stellar explosion so far known (Gamma Ray Bursts) has been discussed in several works. These stellar explosions could have left an imprint in the biological evolution on Earth and in other habitable planets. In this work we calculate the short-term lethality that a GRB would produce in the aquatic primary producers on Earth. This effect on life appears as a result of ultraviolet (UV) re-transmission in the atmosphere of a fraction of the gamma energy, resulting in an intense UV flash capable of penetrating ~ tens of meters in the water column in the ocean. We focus on the action of the UV flash on phytoplankton, as they are the main contributors to global aquatic primary productivity. Our results suggest that the UV flash could cause an hemispheric reduction of phytoplankton biomass in the upper mixed layer of the World Ocean of around 10%, but this figure can reach up to 25 % for radiation-sensitive picoplankton species, and/or in conditions in which DNA repair mechanisms are inhibited.
physics.bio-ph
physics
Ultraviolet perturbations Research article Intense ultraviolet perturbations on aquatic primary producers Mayrene Guimarais1, Rolando Cárdenas2 and J.E. Horvath3 [1] {Marine Ecology Group, Center for Research of Coastal Ecosystems, Cayo Coco, Ciego de Avila, Cuba . Phone 53 33 301104 ext 117 e-mail: [email protected]} [2] {Department of Physics, Universidad Central de Las Villas, Santa Clara, Cuba. Phone 53 42 281109 Fax 53 42 281130 e-mail: [email protected]} [3] {Departmento de Astronomia, Instituto de Astronomia, Geofísica e Ciências Atmosféricas, Universidade de São Paulo, Brazil. Phone (11) 3091 2710 e-mail: [email protected]} Abstract During the last decade, the hypothesis that one or more biodiversity drops in the Phanerozoic eon, evident in the geological record, might have been caused by the most powerful kind of stellar explosion so far known (Gamma Ray Bursts) has been discussed in several works. These stellar explosions could have left an imprint in the biological evolution on Earth and in other habitable planets. In this work we calculate the short-term lethality that a GRB would produce in the aquatic primary producers on Earth. This effect on life appears as a result of ultraviolet (UV) re-transmission in the atmosphere of a fraction of the gamma energy, resulting in an intense UV flash capable of penetrating ~ tens of meters in the water column in the ocean. We focus on the action of the UV flash on phytoplankton, as they are the main contributors to global aquatic primary productivity. Our results suggest that the UV flash could cause an hemispheric reduction of phytoplankton biomass in the upper mixed layer of the World Ocean of around 10%, but this figure can reach up to 25 % for radiation-sensitive picoplankton species, and/or in conditions in which DNA repair mechanisms are inhibited. 1 Introduction The ultraviolet shielding problem has attracted the attention of the biologists for a long time. There are many evidences indicating the lack of an atmospheric ultraviolet shield during the Archean eon, and as a consequence the possible onset of a harsh photobiological regime in the planet’s surface, listed among the reasons of why continents might not have been conquered by life until the appearance of the ozone layer. Thus, photobiology stands as one of the drivers of biological evolution in our planet. Radiations in general have the dual role of sterilizing non-resistant species and favouring speciation of the surviving ones, due to DNA mutations and expected availability of ecological niches. Therefore, radiation bursts are plausible hypotheses to explain biodiversity drops and its subsequent increases, as in the case of the Cambrian explosion and Phanerozoic extinctions. One of the natural mechanisms capable of delivering on Earth radiation bursts intense enough are stellar explosions, provided the explosion occurs not too far (Thorsett 1995, Scalo and Wheeler 2000, Galante and Horvath 2007; Martin et al 2009). These explosions typically occur in very massive stars, progenitors of the so-called gamma ray bursts (GRB’s) and associated supernovae. More specifically, it has been suggested that the Ordovician-Silurian mass extinction was caused by a GRB (Melott and Thomas 2009; Melott et al 2004) and that the minor marine extinction in the Pleistocene-Pliocene of tropical bivalves was the consequence of a nearby (ordinary?) supernova (Benítez, Maiz-Apellaniz and Canelles 2002). It is worth noting that the last hypothesis is receiving support by isotopic anomalies in marine sediments of the Pleistocene, namely of the same epoch in which our Solar System neared the Scorpius-Centaurus association of massive stars (Fields, Hochmuth and Ellis 2005). Currently, there is at least one star close enough to Earth as to be considered dangerous because of its potential explosion as a supernova and even the emission of a GRB. It is WR 104, located just 8000 light years away from us (Tuthill et al 2008), although recent spectroscopic measurements suggest a pitch of the gamma beam ~ 30-40 degrees from Earth, thus probably preventing the incidence on our atmosphere. In general, statistical estimates suggest a very low incidence probability, mainly because of a small solid angle gamma emission (Mézáros 2001), but there might be undetected binary systems close enough to be problematic. An earlier work (Galante and Horvath 2007) has compiled and discussed the several effects that a stellar explosion can cause on Earth’s atmosphere and biosphere. The best studied is the depletion of the ozone layer, allowing more solar UV to reach the planet’s ground during several years. In this work, however, we focus on another important short-term effect: the brief and immediate UV-flash reaching the ground as a result of reprocessing the gamma energy in the atmosphere. A closely related phenomenon is also known form solar observations:, arguably our Sun has the potential of sporadic flares intense enough to cause ecological catastrophes, but so far none of them has been recorded and confirmed. However, it is clear that even modest depletions of the ozone layer can significantly influence terrestrial biota, through the enhanced solar UV flux reaching every day the surface of Earth. That is why since the 1970’s much attention has been given to the current depletion of the ozone layer, mainly in the context of the anthropogenic global warming. Actually, the ozone hole over Antarctica and a potential future one over Europe have been defined as tipping points of our planet (Lenton et al. 2008; IPCC 2007). In this work we use tools developed to model biological effects of current ozone depletion in order to do some estimates of the immediate lethality that a stellar explosion or an unusually intense solar flare would cause on phytoplankton. These are the main primary producers and the starting point of the food web in central ocean basins, and are also important in coastal and freshwater ecosystems. Astrophysical calculations based on star formation rate suggest that in the last few billion years each planet in our galaxy would have been affected by a GRB (Scalo and Wheeler 2004). We thus focus in the short-term lethality that would produce on Earth the so-called “typical” GRB of the last billion years: a burst arriving from 3000-6000 light years away and delivering ~ 100 kJ/m2 of gamma energy at the top of the atmosphere. The main short- term difference between stellar explosions and solar flares will be in the specific ultraviolet spectrum deposited at ground, provided total energy is equal. The photobiological methods used to estimate biological damage are, though, the same for each case. 2 Materials and methods 2.1 - The interaction of stellar gamma radiation with the atmosphere The interaction of the gamma burst with the atmosphere has been considered and we adopted the results of Martin et al. (2009) in this work. In first place, the fraction of gamma photons directly reaching the ground is negligible, because of the large Compton cross-section with electrons from the molecules of the atmosphere. The free electrons would then excite other molecules, causing a rich aurora-like spectrum, which reaches the sea level. The ultraviolet fraction of this spectrum (termed the UV-flash) is a major danger for life (Galante and Horvath 2007). The duration of the UV-flash would be the same of the GRB (around 10 seconds), with a high intensity and even including the very deleterious UV-C band in the wavelength range . The interactions of these UV flash photons in water and their efficiency for phytoplankton damage is our concern in this work. 2.2 - Radiative transfer in water and effective doses We considered an average ocean albedo of 6.6 % for zenithal angles not greater than 70 degrees, as reported in Cockell (2000). This was employed to calculate the GRB-UV spectrum just below the ocean surface . We used the classification of optical ocean water types originally presented in Jerlov (1951, 1964, 1976). Consequently, we use the attenuation coefficients K() of UVR in oceanic water types I, II and III as in (Peñate et al 2010). These optical water types can roughly be identified as oligotrophic, mesotrophic and eutrophic, respectively. However, we also included the intermediate types IA and IB. We utilized biological action spectrum for DNA damage e() following Cockell (2000). Then, the (effective) biological irradiances or dose rates E*(z) at depth z follow from : (1) The (effective) biological fluences or doses F*(z) are given by (2) where t is the exposure time to UVR. We also consider that, just before the UV flash, phytoplankton were homogeneously distributed in the upper mixed layer (UML) of the ocean, due to the mixing action of Langmuir currents and related circulation patterns. The depth of UML depends on ocean surface winds and other factors, but after averaging its value for 13 locations (Agustí and Llabrés 2007) we consider it to be 30 meters, quite a typical value. 2.3 – The estimation of induced lethality nm2802600,0EE*(z)eE0(,0)eK()zdtzEzF)(*)(* Experiments with phytoplankton stressed by UVR are typically done exposing them to solar radiation during several hours. This is not a scenario close enough to the one we study, given the low intensity of solar UV, as compared to the GRB UV-flash. Therefore, as done by some of us in (Galante and Horvath 2007), we chose the results of (Gascon et al 1995). These authors intensely irradiated a representative set of bacteria with a “hard” wavelength ( ) of the UV-C band. We believe that the more radiation-sensitive phytoplankton would behave as Escherichia coli, the intermediate as the aquatic bacterium Rhodobacter sphaeroides (wild type and phototrophically grown strain), while the toughest would parallel the soil bacterium Rhizobium meliloti. We also analyzed the case in which repair mechanisms would be inhibited: very cold waters or a night-time UV-flash (because at night cell division is synchronized in oceanic phytoplankton (Agawin and Agustí 2005), making them much more radiation sensitive). To account for this last scenario we use the data in (Gascon et al 1995) for strains in which repair is inhibited due to the lack of recA gene. These data are namely available for the two extremes of our “survival band” (E. coli and R. meliloti). Strains having above gene are denoted recA+, while the absence is indicated by recA-. Starting with the classical model for survival curves of irradiated cells, (3), where is the survival fraction, is the slope and is the dose or fluence, we have introduced some significant refinements. Since the effective biological dose calculated from eqs. (1) and (2) needs to be employed, we propose a refined survival model: (4) where is the surviving fraction at depth , is the new (effective biological) slope, and is the effective biological dose or fluence at depth . The slopes are a measure of the radiosensitivity of the species. We determine them considering that the reported doses in Gascon et al (1995) follow the simple formula: 254nmFeSSFF*)(**zFezSSzzF*zzF (5) Dividing eq. (2) by eq. (5) we obtain: (6) Both F and E are given in Gascon et al (1995), while E* was determined by Cockell (2000) by biologically weighting it, so above equation allows the calculation of F* for each species. We then obtained the biological effective dose for which 10% of the cells survive (F* 10) using the F10 values for each species reported in Gascon et al (1995), and finally found the new slope . 3 Results 3.1 Radiation transfer and effective doses in the ocean The set of attenuation coefficients used is shown in Fig. 1. Notice that in the wavelength range used (260-350nm), the optical quality of types I, IA and IB is not very different. Fig. 1 Attenuation coefficients for the five optical ocean water types in the wavelength range used. FEtFFEE00.20.40.60.811.2250270290310330350370Wavelength; nmK; 1/mIIAIBIIIII The effective biological doses F* delivered in above water types are plotted in Fig. 2, as a function of depth z. Again waters of types I, IA and IB follow a more or less similar behaviour. Fig. 2 Effective biological doses vs. depth for all ocean optical water types 3.2 The estimation of induced lethality In Figs. 3-7 we show the surviving fraction of cells after the GRB UV-flash strikes, for the five optical ocean water types. Fig. 3 Surviving fraction of cells after the GRB UV-flash strikes, for the case of water type I, the more oligotrophic and clear. 05010015020025030035040045050005101520253035z; mF*; J/m2IIAIBIIIII02040608010012005101520253035z; mS; %R. meliloti recA+R. sphaeroidesR. meliloti recA-E. coli recA+E. coli recA- Fig. 4 Surviving fraction of cells after the GRB UV-flash strikes, for the case of water type IA. Fig. 5 Surviving fraction of cells after the GRB UV-flash strikes, for the case of water type IB. Fig. 6 Surviving fraction of cells after the GRB UV-flash strikes, for the case of water type II. 02040608010012005101520253035z; mS; %R. meliloti recA+R. sphaeroidesR. meliloti recA-E. coli recA+E. coli recA-02040608010012005101520253035z; mS; %R. meliloti recA+R. sphaeroidesR. meliloti recA-E. coli recA+E. coli recA-02040608010012005101520253035z; mS; %R. meliloti recA+R. sphaeroidesR. meliloti recA-E. coli recA+E. coli recA- Fig. 7 Surviving fraction of cells after the GRB UV-flash strikes, for the case of water type III. In the Table 1 below we present the surviving fractions in the upper mixed layer of the ocean (30 meters depth) resulting from the above calculations. Total biomass reduction (%) in the mixed layer (30 meters depth) for oceanic optical water types Scenario Species Good R. meliloti repair recA+ I 10,4 IA 8,8 IB 8,8 R. 12,8 10,8 9,1 sphaeroides II 4,2 5,2 E. coli recA+ 20,3 17,2 14,5 8,2 Bad R. meliloti 18,6 15,8 13,2 7,5 III 2,7 3,3 5,0 4,6 repair recA- (cold water or E. coli recA- 57,1 48,6 40,7 22,8 13,3 night UV- flash) Table 1 Total reduction of biomass in the upper mixed layer of the ocean 02040608010012005101520253035z; mS; %R. meliloti recA+R. sphaeroidesR. meliloti recA-E. coli recA+E. coli recA- 3.3 The role of UV-C An interesting feature of the action of the GRB is the presence of irradiances at ground level in the very deleterious UV-C band. In this case we computed non-negligible values between 260 and 280nm. This band is not usually considered because it is totally absorbed by the atmosphere, but the aurora-like spectrum provoked by the GRB includes these wavelengths, emitted in the atmosphere at altitudes low enough as to reach ground. Despite its relatively low intensity, this band per se would account for a significant lethality, as shown in Table 2. Total biomass reduction (%) in the mixed layer (30 meters depth) for oceanic optical water types Scenario Species Good R. meliloti repair recA+ I 2,0 IA 1,7 IB 1,4 R. 2,6 2,2 1,8 sphaeroides E. coli recA+ 4,7 3,9 3,3 Bad R. meliloti 4,2 3,5 3,0 II 0,9 1,1 2,0 1,8 III 0,6 0,8 1,3 1,2 repair recA- (cold water or E. coli recA- 27,0 22,7 19,1 10,7 6,4 night UV- flash) Table 1 Total reduction of biomass in the upper mixed layer of the ocean had the UV- flash contained only the UV-C band 4 Conclusions Most areas of modern ocean basins are oligotrophic (water types I, IA and IB), thus from Table 1 we might expect a lethality of ~10% from a gamma-ray illumination in good repair scenarios, assuming that most species of phytoplankton would behave similarly to the aquatic bacterium R. sphaeroides. However, the cells of some species of picoplankton are so small (diameter around 0,6 m), that it is unlikely that they can host an elaborated DNA repair machinery. An outstanding example is the genus Prochlorococcus. Due to its wide distribution and small size, Prochlorococcus spp. have been termed the most abundant organisms on Earth (Partensky, Hess and Vaulot 1999). In fact, they account for an estimated 20% of the oxygen released to the Earth's atmosphere through the photosynthesis process, and are at the very base of the ocean food assemblage. Given their poor repair capabilities, lethality of species of this genus could reach 25% even in warm waters, and higher at night or in cold waters, where the repair mechanisms are additionally inhibited (Table 1). Also, a night flash would affect several organisms commonly found in deep waters during daylight time. However, it is true that the brief flash of ~ ten seconds would only influence one hemisphere of the planet, the one facing the gamma beam, a fact that attenuates the size of the damage. It is also important to note that phytoplankton living beneath the mixed layer at the moment of the UV-flash would not be affected, even in oligotrophic waters I, IA and IB (Figs. 3-7). In Peñate et al (2010) a total (100%) inhibition of photosynthesis down to 75 meters in water type I was estimated. Comparison with Table 1 then leads us to state that this is mainly due to damages of the photosynthetic apparatus, and not to the other possible cause (induced lethality due to DNA damage). Therefore, estimation of the velocity of recovery to normal population numbers becomes very complicated, depending on: the repair mechanisms of both the photosynthetic apparatus and DNA, the ocean circulation bringing unaffected individuals from below the mixed layer and from the unaffected hemisphere, the extent of the depletion of the ozone layer (which would last a decade or so), the potential climate change (we refer interested readers to Thomas et al 2005 and Galante and Horvath 2007 for a compilation of many potential effects). Aquatic food webs having a strong dependence on phytoplankton might be very affected and it turns out interesting to evaluate the response of the other primary producers (macroalgae and seagrasses). The darker the water the lesser the affectation, therefore in shallow waters the more protected ecosystems would be some very eutrophic coastal and inland waters. References Agawin, N.S.R. and Agustí, S.: Prochlorococcus and Synechococcus cells in the Central Atlantic ocean: distribution, growth and mortality grazing rates. Vie et Milieu 55: 165-175, 2005. Agustí, S. and Llabrés, M.: Solar Radiation-induced Mortality of Marine Pico- phytoplankton in the Oligotrophic Ocean. Photochemistry and Photobiology 83: 793– 801, 2007. Benítez, N., Maiz-Apellaniz J. and Canelles, M.: Evidence for Nearby Supernova Explosions, Phys.Rev.Lett. 88, 081101, 2002. Cockell, C.: Ultraviolet radiation and the photobiology of Earth’s early oceans. Orig. Life Evol. Biospheres 30, 467–499, 2000. Fields, B., Hochmuth, K. and Ellis, J.: Deep-Ocean Crusts as Telescopes: Using Live Radioisotopes to Probe Supernova Nucleosynthesis. Astrophys.J.621, 902-907, 2005. Gascon, J., Oubina, A., Perez-Lezaun, A. and Urmeneta, J.: Sensitivity of selected bacterial species to UV radiation, Current Microbiol. 30, 177–182, 1995. Galante, D. and Horvath, J.E.: Biological Effects of Gamma-Ray Bursts: distances for severe damage on the biota. International Journal of Astrobiology, 6, 19-26 , 2007. IPCC: Intergovernmental Panel on Climate Change: Climate Change 2007 – The Physical Science Basis, Cambridge University Press, 2007. Jerlov, N. G.: Optical Studies of Ocean Water. Report of Swedish Deep-Sea Expeditions 3, 73–97 , 1951. Jerlov, N. G.: Optical Classification of Ocean Water. In Physical Aspects of Light in the Sea. (Honolulu:University of Hawaii Press), 45–49 , 1964. Jerlov, N. G.: Applied Optics, Elsevier Scientific Publishing Company, Amsterdam, 1976. Lenton, T. M., H. Held, E. Kriegler, J. W. Hall, W. Lucht, S. Rahmstorf and H. J. Schellnhuber : Tipping elements in the Earth’s climate system, Proceedings of the National Academy of Sciences USA 105(6), 1786–1793, 2008. Martin, O., Galante, D., Cardenas, R. and Horvath, J.E, : Short-term effects of gamma ray bursts on Earth. Astrophys Space Sci, 321, 161–167, 2009. DOI 10.1007/s10509- 009-0037-3 Melott, A. and Thomas, B.: Late Ordovician geographic patterns of extinction compared with simulations of astrophysical ionizing radiation damage, Paleobiology 35, 311-320, 2009. Melott, A., Lieberman, B., Laird, C., Martin, L., Medvedev, M., Thomas, B., Cannizzo, J., Gehrels, N. and Jackman, C.: Did a gamma-ray burst initiate the late Ordovician mass extinction? Int.J.Astrobiol.3, 55, 2004. Mézáros, P.: Gamma-ray bursts. Science 291, 79-83, 2001. Partensky, F., Hess, W., Vaulot, D.: Prochlorococcus, a Marine Photosynthetic Prokaryote of Global Significance, Microbiology and Molecular Biology Reviews, 63, 106, 1999. Peñate, L., Martín, O., Cardenas, R. and Agustí, S.: Short-term effects of Gamma Ray Bursts on oceanic photosynthesis http://arxiv.org/abs/1007.2879 Thomas, B., Melott, A., Jackman, C., Laird, C., Medvedev, M., Stolarski, R., Gehrels, N., Cannizzo, J., Hogan, D. and Ejzak, L.: Gamma-Ray Bursts and the Earth: Exploration of Atmospheric, Biological, Climatic and Biogeochemical Effects. Astrophys.J. 634, 509-533 , 2005. Tuthill, P., Monnier, J., Lawrence, N., Danchi, W., Owocki, S. and Gayley, G.: The Prototype Colliding-Wind Pinwheel WR 104, Astrophysical J. 675, 698-705, 2008.
1608.06590
1
1608
2016-08-23T17:41:47
Identification of Bifurcations from Observations of Noisy Biological Oscillators
[ "physics.bio-ph", "q-bio.CB", "q-bio.NC" ]
Hair bundles are biological oscillators that actively transduce mechanical stimuli into electrical signals in the auditory, vestibular, and lateral-line systems of vertebrates. A bundle's function can be explained in part by its operation near a particular type of bifurcation, a qualitative change in behavior. By operating near different varieties of bifurcation, the bundle responds best to disparate classes of stimuli. We show how to determine the identity of and proximity to distinct bifurcations despite the presence of substantial environmental noise.
physics.bio-ph
physics
Please cite this article as: Salvi J. D., Ó Maoiléidigh D., and Hudspeth A. J. (2016) Identification of bifurcations from observations of noisy biological oscillators. Biophysical Journal 111(4):798-812. http://dx.doi.org/10.1016/j.bpj.2016.07.027 Read the complete article online: http://www.cell.com/biophysj/fulltext/S0006-3495(16)30596-3 Supporting Material can be downloaded online from: http://www.cell.com/cms/attachment/2062690836/2065160716/mmc1.pdf 2 Identification of bifurcations from observations of noisy biological oscillators Joshua D. Salvi, Dáibhid Ó Maoiléidigh, and A. J. Hudspeth† Howard Hughes Medical Institute and Laboratory of Sensory Neuroscience, The Rockefeller University, 1230 York Avenue, New York, New York, 10065, USA Running Title: Identification of noisy bifurcations † Correspondence: [email protected] 3 Abstract Hair bundles are biological oscillators that actively transduce mechanical stimuli into electrical signals in the auditory, vestibular, and lateral-line systems of vertebrates. A bundle's function can be explained in part by its operation near a particular type of bifurcation, a qualitative change in behavior. By operating near different varieties of bifurcation, the bundle responds best to disparate classes of stimuli. We show how to determine the identity of and proximity to distinct bifurcations despite the presence of substantial environmental noise. Using an improved mechanical-load clamp to coerce a hair bundle to traverse different bifurcations, we find that a bundle operates within at least two functional regimes. When coupled to a high-stiffness load, a bundle functions near a supercritical Hopf bifurcation, in which case it responds best to sinusoidal stimuli such as those detected by an auditory organ. When the load stiffness is low, a bundle instead resides close to a subcritical Hopf bifurcation and achieves a graded frequency response-a continuous change in the rate but not the amplitude of spiking in response to changes in the offset force-a behavior useful in a vestibular organ. The mechanical load in vivo might therefore control a hair bundle's responsiveness for effective operation in a particular receptor organ. Our results provide direct experimental evidence for the existence of distinct bifurcations associated with a noisy biological oscillator and demonstrate a general strategy for bifurcation analysis based on observations of any noisy system. 4 Introduction A bifurcation occurs when a quantitative change in the value of some parameter-a control parameter-induces a qualitative change in the behavior of a system. Bifurcations are often encountered in theoretical and experimental systems in physics, chemistry, biology, medicine, economics, and climatology (1-4). All systems operating near a given type of bifurcation exhibit similar dynamics (5), and systems operating near different bifurcations can exhibit distinct behaviors. Identifying a bifurcation therefore reveals generic properties of a complex system and permits prediction of that system's function when it operates near the bifurcation. The activity of spiking neurons, for example, can be segregated into at least two classes of excitability with distinct patterns in the frequency of spiking that are attributed to operation near different types of bifurcation (6-8). Class 1 neurons display a continuous change in spike frequency in response to changes in stimulus current, a behavior commonly ascribed to their operation near a saddle-node on invariant cycle (SNIC) bifurcation (9). Class 2 neurons, by contrast, exhibit a discontinuous jump in frequency as a function of the applied current, a behavior consistent with their crossing a Hopf bifurcation (10). By identifying the type of bifurcation, one can assess generic features of neuronal excitability in different neuronal populations. Bifurcation analysis is usually conducted on purely mathematical stereotypes or on simplified representations of real-world systems for which the effects of noise are small. In certain experimentally accessible systems, however, noise plays a significant and irreducible role in shaping the dynamics. Among these systems is the hair cell, the sensory receptor of the auditory, vestibular, and lateral-line sensory systems of vertebrate animals (11). A hair cell detects mechanical signals derived from sounds, accelerations, and water movements, transducing them into electrical signals that are transmitted to the brain. This detection is achieved through the motion of a mechanosensitive organelle-the hair bundle-that projects 5 from the hair cell's apical surface and transduces mechanical input into electrical output in the cell body. For small-magnitude stimuli the hair bundle operates in an environment dominated by noise. In fact, the sensitivity of our hearing is limited by the clattering of air molecules against the eardrum and the rattling of water molecules within the cochlea (12, 13). Thermal noise also plays an important role in setting the sensitivity of vestibular organs (14). Auditory organs deal with noise by employing an active process, a metabolically powered mechanism that enhances their sensitivity (15). Vestibular systems might also employ the active process to improve their effectiveness. The effects of the active process depend on the values of a sensory organ's parameters (16). For example, it has been hypothesized that auditory hair bundles achieve enhanced frequency selectivity and a broadened dynamic range in response to periodic stimuli when they are poised near the onset of spontaneous oscillation, that is, when they operate close to a supercritical Hopf bifurcation (17-19). Observations of individual bundles support this hypothesis (20). Hair bundles have also been observed to attain a graded frequency response with changes in static offsets, which could occur if hair bundles operate near a SNIC bifurcation or-as we argue below-a subcritical Hopf bifurcation (20-22). We propose that this graded response allows hair bundles to function as static force detectors, which would be of use in a vestibular system. By operating near particular bifurcations, a mechanically active hair bundle can therefore be useful in different mechanosensory systems. A bundle's state diagram depicts its behavior for different operating points defined by the values of the system's control parameters (16, 20). These control parameters include the mechanical loads imposed on individual hair bundles by accessory structures in vivo, such as the tectorial membrane coupled to the bundles of outer hair cells in the mammalian cochlea or the otolithic membrane coupled to hair cells of the utriculus and sacculus. Each structure loads the bundle with a mass, a drag, a stiffness, and an offset force; the combination allows a bundle to operate near the type of bifurcation that is suitable for its sensory role. 6 Although biophysical experiments have identified several of the mechanisms that underlie the active process (15), the noisy environment of a hair bundle complicates the identification of the relevant bifurcations. In particular, noise blurs the boundaries between the dynamical regimes in the bundle's state diagram and conceals the characteristics of various bifurcations. There are several types of bifurcation in which a quiescent system becomes self- oscillatory. In the absence of noise, a system crossing either a SNIC or a Hopf bifurcation exhibits respectively a continuous rise or a discontinuous jump in the frequency of spontaneous oscillation (5-7). Changes in the amplitude of oscillation with the adjustment of a parameter further distinguish two types of Hopf bifurcation: a gradual rise in amplitude corresponds to a supercritical Hopf bifurcation and a discontinuous jump signals a subcritical Hopf bifurcation (7). In the presence of noise, however, sharp transitions in the amplitude and frequency of spontaneous oscillation are blunted and become difficult to differentiate from their gradual counterparts, complicating both the localization and the identification of bifurcations. We have overcome this challenge by employing a battery of statistical tests to locate and identify bifurcations from experimental observations of a noisy system. Deterministic bifurcations, those defined in the absence of noise, occur at well-specified parameter values. For example, a deterministic system possessing a supercritical Hopf bifurcation is quiescent on one side of the bifurcation and oscillatory on the other. Although the location of this bifurcation can only be estimated from observations, we can nonetheless identify its quiescent and oscillatory sides to define the bifurcation empirically. By analogy with the deterministic case, we have defined the location of an empirical bifurcation for a noisy system as the parameter value at which spontaneous oscillations can be reliably detected from observations. The quiescent side of this empirical bifurcation covers the range of parameter values for which oscillations cannot be detected with sufficient statistical certainty. In this manner we have delineated the oscillatory and quiescent regions near supercritical Hopf, subcritical Hopf, and SNIC bifurcations from noisy data. 7 We have explored various metrics that capture the distinguishing features of systems crossing different bifurcations. To do so, we have compared simulations of noisy systems near various bifurcations with experimental observations of noisy hair-bundle motion. Based on these comparisons, we have ascertained which metrics best characterize the identities and locations of bifurcations near which hair bundles operate. By employing these metrics to assess the behavior of individual hair bundles, we have established that a hair bundle can operate near at least two types of bifurcation. We have then associated the bundle's operation near these bifurcations with the functions of auditory and vestibular receptor organs. Our methodology not only furthers an understanding of hair-bundle dynamics but also provides a general strategy for analyzing empirical observations of noisy dynamical systems. Materials and Methods All experiments were performed on spontaneously active hair bundles from the saccular maculae of adult bullfrogs, Rana catesbeiana (20). Supporting Material. We have included detailed Supporting Material to accompany this paper. Although the entirety of the document is not required for a general understanding of our findings, we prepared the Supporting Material so that it can serve as a self-contained guide for those who wish to apply our approach to other systems. Mechanical-load clamp. We employed a feedback system with a real-time interface to control the load stiffness and constant force applied to individual hair bundles (23). Using a glass fiber as a model hair bundle, we confirmed that the clamp achieves this performance with high precision and accuracy (Fig. S1). For a complete description of the mechanical-load clamp, see Supporting Material Section A. 8 State-diagram mapping. To calculate the amplitude and frequency of hair-bundle motion in response to various combinations of load stiffness and constant force, we employed both described methods (20) and a peak-detection algorithm (24). Operating points for which the bundle's position histogram displayed at least two clear maxima, as signaled by a Hartigans' dip statistic that exceeded 0.01 with p < 10-3, were classified as oscillatory (25). All other operating points were deemed quiescent. The Hartigans' dip statistic is described in more detail below, and an expanded explanation of state-diagram calculations can be found in Supporting Material Section B. Time-series analysis. We employed a battery of statistical tests to assess the behavior of experimentally observed hair bundles and of simulated systems. We included simple tests that limited the number of manually selected parameters. Our approach requires only noisy time- series data and does not necessitate stimulation. These features enhance the versatility of the method, permitting its use in systems for which stimulation is difficult or impossible. For a complete description of these metrics, see Supporting Material Section C. A bifurcation occurs when a system undergoes a qualitative change in behavior. For example, the system may transition from a domain of quiescence to one of spontaneous oscillation. We first sought to estimate the location of a bifurcation by detecting the onset of spontaneous oscillations across a range of control-parameter values. A system that oscillates spontaneously displays a position distribution with more than one peak (20, 26, 27). Although previous studies measured the distance between peaks in the distribution to describe this phenomenon (27, 28), we instead employed Hartigans' dip statistic to measure the modality of a distribution (25). Not only is the dip statistic an inferential measure of modality with an associated p-value, it is also less biased by skew and sample size relative to other metrics (29). Whereas a small value of the dip statistic corresponds to a position distribution with one peak, a large value reflects a distribution that has more than one peak. We therefore defined a position 9 distribution possessing a large value of the dip statistic with a correspondingly small p-value as an indicator of spontaneous oscillation, and a small value of the dip statistic as an indicator of quiescence. The behavior of a system is typically recorded using only a single observable, such as a bundle's position in time. However, in many instances a system is described by more than one variable and the system's dynamics occupies a space of many variables called a phase space. A dynamical trajectory in phase space possesses information that is not apparent from a single- variable time series. We therefore wished to reconstruct a two-dimensional phase-space representation of hair-bundle dynamics based only on observations of its displacement. Because many methods of phase-space reconstruction require the manual selection or empirical estimation of several parameter values (30-33), we focused instead on a simpler alternative. The Hilbert transform of a signal is the imaginary part of its analytical representation (34). The joint probability distribution of a bundle's real-valued position and the Hilbert transform of that position-an analytic distribution-is a distribution over a system's phase space that can be calculated without the need for complicated analysis and challenging parameter selection. The analytic distribution appears ring-shaped when a system displays limit-cycle oscillations, which correspond to stable, closed-loop trajectories in phase space. Near quiescent fixed points the distribution instead exhibits an enhanced, disk-shaped density. This representation is similar to that used to analyze spontaneous otoacoustic emissions (35). We next estimated the frequency, amplitude, and regularity of spontaneous oscillations with a peak-detection algorithm that allows the detection in a time series of all peaks and troughs that cross defined thresholds (24). The method permits calculation of a system's amplitude and frequency of oscillation for time series with substantial noise, short durations, and non-sinusoidal waveforms. For each peak-detection threshold, the frequency is defined as the inverse of the mean time interval between peaks, and the amplitude is calculated as half of the mean difference in position between each peak and trough. 10 We did not employ spectral methods to determine the location or identity of a bifurcation. In the presence of noise, both a quiescent resonant system and a limit-cycle system exhibit peaks in their Fourier spectra (36). A qualitative change in spectra as this noisy system crosses a bifurcation does not exist. In contrast, a system's position distribution exhibits a qualitative change when the system crosses a bifurcation. Moreover, because many of our signals were subjected to substantial noise and displayed non-sinusoidal waveforms, the Fourier frequency and amplitude of oscillation often fluctuated considerably as the value of a control parameter was changed, and consequently this approach did not yield a clear estimate of a bifurcation's location that was consistent across the systems we studied (Figs. S6-S8). The peak-detection algorithm also performed more reliably than spectral analysis in estimating the amplitude and frequency of noise-induced spikes (Fig. S9). As a system transitions from a regime dominated by limit-cycle oscillations to one dominated by noise, that system's spontaneous motion becomes more irregular. The distribution of interpeak time intervals, the times between neighboring peaks, correspondingly broadens. We therefore assessed the regularity of a system's oscillations by calculating the coefficient of variation, defined as the ratio of this distribution's standard deviation to its mean. As a system's oscillations become increasingly irregular, the metric grows and thus traces the transition from motion governed by limit-cycle oscillations to that dominated by noise. The regularity of oscillations determined using the coefficient of variation implies that multiple peaks in the position histogram stem from limit-cycle oscillations rather than from noise-induced switching between stable states. For example, a system crossing a saddle-node, or fold, bifurcation may possess a stable state on one side of the bifurcation and two stable states on the other. This system can stochastically switch between two states in the bistable regime. We found that the coefficient of variation for a noisy system transitioning from monostability to bistability is large for all control-parameter values, in contrast with the smaller coefficients of variation displayed by systems displaying limit-cycle oscillations (Fig. S12). 11 Finally, we estimated the location of a bifurcation using an information-theoretical metric. Upon crossing certain bifurcations, a system exhibits limit-cycle oscillations that appear ring-shaped in that system's analytic distribution. To quantify this phenomenon, we estimated the mutual information between the real and imaginary parts of the system's analytic signal, which we entitle the analytic information. The analytic information resembles the time-delayed mutual information, a nonlinear measure of temporal correlation in a signal (37). As opposed to a shift of the signal in time, however, the analytic information is calculated from a real-valued signal and that signal shifted in phase. In the presence of a limit cycle the analytic information is large; in the absence of a cycle its value is small. For example, the analytic information approaches zero for a narrow-band Gaussian-distributed noise sequence as its length increases (Fig. S10). The information increases as a system crosses a bifurcation and begins to oscillate spontaneously. Together with the dip statistic, the analytic information therefore serves as an indicator of the onset of limit-cycle oscillations. Numerical simulations. A system operating near a bifurcation can be described by the normal form corresponding to the type of bifurcation. The normal form is a standard mathematical expression that captures the generic features of any system operating near that kind of bifurcation. We performed simulations of the normal forms for different bifurcations, in which we included various levels of additive noise for each variable. We simulated the normal forms of the supercritical Hopf, the subcritical Hopf, and the SNIC bifurcations. We also simulated a model of hair-bundle dynamics with additive noise (16). We refer to all these simulations as stochastic simulations throughout the paper. All simulations were conducted in MATLAB (R2014a, 8.3.0.532) with the Euler-Murayama method. Because the simulations of the normal forms and the hair-bundle model have no meaningful dimensions in position and time, the results of the simulations are reported without units. In keeping with the convention in the dynamical-systems literature, we orient the abscissa of every plot such that the quiescent regime occurs to the left and the oscillatory regime to the right of each bifurcation. For a complete description of these simulations, see Supporting 12 Material Sections D and E. Results The hair bundles of various receptor organs confront a variety of physical loads (15). Both modeling and experiments show that adjusting the mechanical load qualitatively changes a hair bundle's behavior (16, 20). Our previous work was limited, however, by the stability and precision of the clamp system used to apply the load on a hair bundle. These limitations impeded determination of the location and identity of bifurcations displayed by a bundle. After improving the mechanical-load clamp to overcome these restrictions (Figs. S1-S3), we have probed the nature of bifurcations more systematically than heretofore. The hair bundle's state diagram Changing the load stiffness and constant force applied to an individual hair bundle yields a map of the bundle's behavior-a state diagram-for combinations of these two control parameters (Fig. 1A). We predicted earlier that a hair bundle oscillates spontaneously within a region bounded by subcritical and supercritical Hopf bifurcations and exhibits bistability in a regime bounded by lines of fold bifurcations (16). Within the oscillatory regime, we also expect the bundle's spontaneous motion to fall in amplitude and rise in frequency with an increase in load stiffness. In the present study, we have explored a bundle's behavior within different locales of its state diagram and identified the types of bifurcation near which the bundle can operate. Using the improved load clamp, we determined the behavior of a bundle as a function of its mechanical load. In agreement with the theoretical expectation, the load stiffness controlled the character of the bundle's spontaneous oscillations (Fig. 1B). When the stiffness was small, the bundle oscillated at low frequency and with a high amplitude and in some cases exhibited 13 mixed-mode oscillations. Increasing the load stiffness caused the bundle's oscillations to increase in frequency and decrease in amplitude until they vanished altogether. A spontaneously oscillating hair bundle displays a distribution of positions with more than one peak, whereas the position distribution of a quiescent bundle has only a single peak (20, 26, 27). We therefore classified the bundle's behavior as either oscillatory or quiescent based on its position histogram. To do so, we employed Hartigans' dip statistic, for which a large value corresponds to a multimodal position distribution generated by a spontaneously oscillating bundle and a small value arises from a unimodal position distribution produced by a quiescent bundle (25). This metric revealed an ovoid oscillatory regime surrounded by a domain of quiescence (Figs. 1C-D and S4). The boundary between these two regimes is expected to correspond to lines of subcritical and supercritical Hopf bifurcations (16, 20). Improvements in data acquisition and analysis allowed us to construct the experimental state diagram with greater stability and a larger signal-to-noise ratio over this range of control parameters than heretofore, eliminating the need for additional statistics to delineate the locus of oscillation (20). The frequency of spontaneous oscillation rose and the amplitude fell with an increase in load stiffness. With an increase in the constant force, however, both the amplitude and the frequency declined (Fig. 1E). These correlations accord with theoretical predictions and prior experimental investigation of the effects of load stiffness on a bundle's behavior (16, 20). Although the experimental state diagram depicted the boundary of and patterns within a regime of spontaneous oscillation, the identity of the bifurcations that define the boundary has not been determined definitively. Because a bundle's operation near a particular bifurcation can in part dictate its mechanosensory function, we developed a strategy for the identification of bifurcations in experimental systems with substantial noise. 14 A supercritical Hopf bifurcation for high stiffness loads We first assessed whether a hair bundle in the high-stiffness regime can operate near a supercritical Hopf bifurcation. To do so, we compared simulations of a noisy system crossing a supercritical Hopf bifurcation, described by the bifurcation's normal-form equation, with experimental observations of a hair bundle subjected to a high load stiffness and increasing values of constant force (Fig. 2). Both the model system and the experimentally observed hair bundle oscillated spontaneously with amplitudes that grew with respectively an increase in the control parameter or a decrease in the bundle's constant force (Figs. 2A, H and S5A). When the amplitude of oscillation exceeded the noise, the position histogram displayed two distinct peaks. The analytic distribution, a two-dimensional representation of the bundle's motion, formed a loop corresponding to a limit cycle. An increase in the control parameter or a decrease in the bundle's constant force caused the diameter of the loop to grow in concert with a rise in the amplitude of oscillation. We determined the location of an empirical bifurcation as the point at which the dip statistic assumed a statistically significant value. Our estimate of the noisy bifurcation's location resided on the oscillatory side of a deterministic supercritical Hopf bifurcation (Fig. 2B). For an active hair bundle, we found a sharp transition between small and large values of the dip statistic, corresponding to a bifurcation defining the boundary between a quiescent and an oscillatory regime (Fig. 2I). Although a system's amplitude and frequency of motion are typically calculated from its Fourier transform, the Fourier amplitude and frequency can be uninformative for determining a bifurcation's location when noise is substantial, when the signals are brief, or when oscillations deviate substantially from pure sinusoids. We illustrate this problem with the Fourier transform both for noisy simulations and for experimental observations of hair-bundle motion (Supporting Material Section C.7). For an oscillator operating far from a bifurcation, spectral analysis performs well. Near a bifurcation and in the presence of noise, however, detection of the 15 bifurcation using the Fourier amplitude and frequency becomes difficult. We show that the amplitude and frequency calculated with the Fourier transform fluctuate considerably as a system's control parameter is changed (Fig. S8). Furthermore, spectral methods do not reliably capture the amplitude and frequency for a system that exhibits noise-induced spikes (Fig. S9). Consequently, this method can fail to detect bifurcations in systems dominated by noise. We therefore characterized the amplitude of spontaneous oscillation using two other metrics. First, we calculated the root-mean-square (RMS) magnitude. For both the simulation and the experimentally observed hair bundle, the RMS magnitude rose gradually as the operating point was moved toward the region of spontaneous oscillation (Fig. 2C, J). However, both the amplitude and the frequency of oscillation can affect the RMS magnitude. For example, constant-amplitude spikes that become less frequent correspond to a declining RMS magnitude even though their amplitude does not change. To circumvent this issue, we additionally determined the amplitude of oscillation from a peak-detection algorithm that found the local maxima (peaks) and minima (troughs) separated by a threshold distance. Our peak-detection algorithm accurately estimated the amplitude and frequency of oscillation (Supplementary Material Section C.7). We defined the amplitude as half of the average distance between the position of a peak and its neighboring trough. As the control parameter grew or the bundle's constant force shrank, the amplitude rose sharply and then more gradually regardless of the peak- detection threshold (Fig. 2D, K). The amplitudes calculated from the peak-detection algorithm were sensitive to the selected peak-detection threshold. For both the simulated and the experimentally observed time series, the amplitude curve shifted toward the oscillatory region as the threshold rose. We then used the same peak-detection algorithm to calculate the frequency of oscillation, which we defined as the inverse of the mean interval between successive peaks. As noted for the amplitude relations, the frequencies calculated with this method were sensitive to changes in the peak-detection threshold. Increasing the value of the peak-detection threshold shifted the 16 frequency curve farther into the oscillatory side of the bifurcation for both the noisy simulations and the experimental observations (Fig. 2E, L). As a system transitions from a regime dominated by large-amplitude oscillations to one dominated by noise, its spontaneous motion becomes increasingly irregular. We therefore quantified the variability of a system's spontaneous motion by the coefficient of variation for the system's interpeak time intervals: large and small coefficients correspond respectively to irregular and regular oscillations. For all peak-detection thresholds, this metric fell as the systems' operating points moved farther into the oscillatory regime, indicating less variation in the interpeak intervals as the system's motion became dominated by limit-cycle oscillations (Fig. 2F, M). Increasing the value of the peak-detection threshold shifted the location at which the coefficient of variation crossed a threshold of 0.5 farther into the oscillatory regime for both a noisy system crossing a supercritical Hopf bifurcation and for an experimentally observed hair bundle. As with the amplitude and frequency of motion, the coefficient of variation depended on the peak-detection threshold. The point at which the coefficient crossed 0.5 moved toward the oscillatory regime with an increase in the value of the peak-detection threshold. A noisy bistable or multistable system exhibits a position histogram with more than one peak and consequently a large value for the dip statistic. The dip statistic alone therefore cannot distinguish limit-cycle oscillations from noise-induced switching between stable states. However, position fluctuations are much less coherent for bistable and multistable systems than for a limit-cycle oscillator. We can therefore use the coefficient of variation to determine whether a system exhibits limit-cycle behavior or whether it displays noised-induced switching between stable states. The coefficient of variation never falls below 0.5 for a bistable or multistable system, but it can approach zero for a limit-cycle oscillator (Supporting Material Section D.6). Finally, we sought to pinpoint the location of a bifurcation and to determine its identity with another metric. The analytic information, defined as the mutual information between the 17 real and imaginary parts of the system's analytic signal, approaches zero for normally distributed noise and grows as limit-cycle oscillations emerge. The analytic information rose gradually with an increase in the control parameter in the simulation and with a decrease in the bundle's constant force in the experiment (Fig. 2G, N). Although the gradual rise in the analytic information failed to identify the exact location of a bifurcation, the trends for both the model and experiment displayed strong similarity. Taken together, the striking agreement between simulations and experimental observations implies that a hair bundle subjected to a large load stiffness can traverse a supercritical Hopf bifurcation as the constant force is changed. To confirm that the high-stiffness boundary between the oscillatory and quiescent regions constitutes a line of supercritical Hopf bifurcations, we subjected both a model hair bundle and an experimentally observed one to a constant force of zero and decreasing values of the load stiffness, a regime for which the model bundle is known to cross a supercritical Hopf bifurcation. We then employed the same battery of tests as before to assess the similarities between the time series of the simulated and experimentally observed hair bundles (Fig. 3). All panels in Figure 3 display the same metrics as the corresponding panels in Figure 2 but for different time series. In agreement with the generic features of a noisy system crossing a supercritical Hopf bifurcation (Fig. 2A), both the simulated and the experimentally observed hair bundle displayed spontaneous oscillations whose amplitude grew with a decrease in stiffness (Figs. 3A, H and S5B). Their analytic distributions also displayed limit cycles that correspondingly increased in diameter. Consistent with that of a noisy system near a supercritical Hopf bifurcation (Fig. 2B), the simulated bundle's position histogram became bimodal on only the oscillatory side of the deterministic bifurcation (Fig. 3B). We observed a clear transition at which simulated and experimentally observed hair bundles became oscillatory (Fig. 3B, I). Both the RMS magnitude (Fig. 3C, J) and the amplitude (Fig. 3D, K) of spontaneous oscillation rose gradually with a decrease in stiffness; this pattern accorded with that of a noisy system crossing a supercritical Hopf bifurcation (Fig. 2C-D). The frequency of oscillation, 18 however, followed a trend that differed from our simulations of generic dynamics near that bifurcation. As the stiffness decreased from its greatest value, the frequency of oscillation first rose for both the simulated and the experimentally observed hair bundle in agreement with a quiescent system's approach to a supercritical Hopf bifurcation (Fig. 3E, L). Further decreases in stiffness, however, caused the frequency to achieve a maximal value and subsequently to decline as the operating point moved beyond the range of influence of the bifurcation. Although this pattern in frequency differs from the generic behavior of a system operating near a supercritical Hopf bifurcation, it accords with the specific behavior predicted for hair-bundle dynamics (Fig. 1A). The coefficient of variation was once again sensitive to changes in the peak-detection threshold, consistent with the generic behavior of a system operating near a supercritical Hopf bifurcation (Fig. 2E-F). For both the simulated and the experimentally observed bundles, the coefficient of variation crossed 0.5 at smaller stiffness values when the value of the peak- detection threshold was larger (Fig. 3E-F, L-M). Finally, the analytic information rose gradually as the systems' operating points advanced into the oscillatory regime (Fig. 3G, N), in agreement with that of a system crossing a supercritical Hopf bifurcation (Fig. 2G, N). All the observations of hair-bundle behavior agreed with stochastic simulations of the normal form of a supercritical Hopf bifurcation and of a model of hair-bundle dynamics. Taken together, these data strongly support the prediction that for an active hair bundle the boundary of oscillation for large stiffnesses is a line of supercritical Hopf bifurcations (16, 20). A subcritical Hopf bifurcation for low stiffness loads We predicted that a hair bundle subjected to a small load stiffness operates near a subcritical Hopf bifurcation (Fig. 1A), whereas an alternate analysis suggested that the bundle approaches a SNIC bifurcation (22, 38). To evaluate these alternatives, we used the methodology discussed above to compare experimental observations of a hair bundle subjected to a small load stiffness to those of noisy systems operating near either a subcritical Hopf or a SNIC bifurcation. All 19 panels in Figures 4 and 5 correspond to the metrics displayed in Figures 2 and 3 but for different sets of simulations and experimental observations. When operating on the oscillatory side of either a subcritical Hopf or a SNIC bifurcation, the noisy systems exhibited large-amplitude oscillations (Fig. 4A, H) that accorded with those of an experimentally observed hair bundle subjected to a small stiffness (Fig. 4O). On the other side of a subcritical Hopf bifurcation, however, this system displayed noise-induced bursts of high- amplitude oscillations (Fig. 4A). A system near a SNIC bifurcation instead exhibited noise- induced spikes with a frequency that fell with a decrease in the control parameter (Fig. 4H). The bursting behavior near a subcritical Hopf bifurcation occurred when a limit cycle coexisted with a stable point at its center, as evidenced by an analytic distribution possessing an enhanced density within a loop (Figs. 4A and S5C). Fluctuations induced transitions between the stable point and the limit cycle, resulting in bursts of noisy oscillations. Bursting occurred only for parameter values within the region of coexistence that fell between the subcritical Hopf bifurcation and a saddle-node of limit cycles bifurcation. The latter bifurcation, at which a stable and an unstable limit cycle collide and annihilate one another, is a consequence of operation near a subcritical Hopf bifurcation. The spiking behavior exhibited by a system near a SNIC bifurcation, by contrast, engendered a locus of high probability along one part of a cycle (Figs. 4H and S5C). The bundle's spontaneous motion most closely resembled the spiking behavior and analytic distributions of a noisy system proximal to a SNIC bifurcation (Figs. 4O and S5C). Whereas the dip statistic indicated noisy oscillations solely on the oscillatory side of a deterministic supercritical Hopf bifurcation (Fig. 2B), it instead evidenced noisy oscillations not only on the deterministically oscillatory side but also within the coexistence region of a deterministic subcritical Hopf bifurcation (Fig. 4B). The dip statistic also detected oscillations on the oscillatory side of a deterministic SNIC bifurcation (Fig. 4I). The dip statistic for an experimentally observed hair bundle clearly delineated the boundary between a quiescent regime 20 and an oscillatory one, again demonstrating the utility of this metric in pinpointing a bifurcation's location (Fig. 4P). All the other metrics were qualitatively indistinguishable between noisy systems operating near a subcritical Hopf or a SNIC bifurcation, with behaviors from both systems resembling those of an experimentally observed hair bundle subjected to a low stiffness. However, these behaviors were distinct from those of a system close to a supercritical Hopf bifurcation (Figs. 2 and 3). In all cases, the RMS magnitude increased abruptly near a bifurcation (Fig. 4C, J, Q). The amplitude rose suddenly before a bifurcation and then grew gradually or remained constant (Fig. 4D, K, R). Unlike the pattern of a system operating near a supercritical Hopf bifurcation, the rise in the frequency of spontaneous oscillation and the fall in the coefficient of variation with an increase in control parameter were insensitive to the value of the peak-detection threshold (Fig. 4E-F, L-M, S-T). Finally, the analytic information rose abruptly near the bifurcation (Fig. 4G, N, U), in contrast with the gradual rise displayed by a system near a supercritical Hopf bifurcation (Fig. 2G). These data together demonstrate that, for a low stiffness, a hair bundle can cross either a subcritical Hopf or a SNIC bifurcation but not a supercritical Hopf bifurcation. We next assessed whether the bursting behavior of a system near a subcritical Hopf bifurcation disqualifies this bifurcation as that exhibited by observed hair bundles at low stiffnesses. Using the same battery of tests, we compared experimental observations of another hair bundle with stochastic simulations of a model bundle that is known to operate near a subcritical Hopf bifurcation (Figs. 1A, 5). In the presence of noise, the simulated bundle exhibited downward excursions resembling the spikes displayed by both of the experimentally observed bundles at low stiffnesses (Figs. 4O, 5A, H, and S5D). Unlike a system displaying bursting near a subcritical Hopf bifurcation (Fig. 4B), the simulated bundle showed noise- induced spikes well beyond the saddle-node of limit cycles bifurcation (Fig. 5B). The bundle displayed a graded frequency response: the spiking frequency fell with an increase in the 21 bundle's constant force. This change was accompanied by a locus of increasing probability in the analytic distribution that rested upon one part of a large-amplitude limit cycle. Spikes may therefore result either from the asymmetric dynamics associated with a SNIC bifurcation or from the specific asymmetry captured by the hair-bundle model near a subcritical Hopf bifurcation, but lacking in the normal form of a subcritical Hopf bifurcation. All metrics showed strong qualitative agreement between the model bundle operating near a subcritical Hopf bifurcation, a SNIC bifurcation, and both experimentally observed hair bundles (Figs. 4 and 5). The changes in the metrics with control parameter were once again distinct from those of a noisy system traversing a supercritical Hopf bifurcation. Discussion A hair bundle's function can be dictated in part by its operation near a particular bifurcation. We have identified the types and locations of bifurcations from experimental observations of noisy hair bundles. By employing several metrics to compare models with experimental observations, we have analyzed the bifurcation structure of experimentally observed hair bundles operating in a noisy environment and confirmed the predictions of a qualitative model of hair-bundle dynamics. Our model makes no assumptions about the temporal or spatial scales of a bundle's dynamics and requires only two properties of hair bundles: their nonlinear stiffness and adaptation to stimulation (15, 39-42). Any actual bundle or hair-bundle model possessing these features exhibits the same state diagram as our model and thus qualitatively similar responses to stimulation (16). Moreover, we previously used a quantitative model of hair-bundle mechanics with physical parameters plausible for the mammalian cochlea to demonstrate that hair-bundle activity is likely essential for the mammalian auditory system to achieve great sensitivity and sharp frequency selectivity in response to high-frequency periodic stimuli (43). 22 Guided by the qualitative model, we found that a single hair bundle can operate near more than one type of bifurcation, depending on its mechanical load. Although it can be argued that noise introduces new bifurcations and changes the character of existing ones (44), our observations accord well with simulations of systems crossing bifurcations in the presence of noise. Proximity to a bifurcation To understand certain dynamical systems one must determine whether they can operate near bifurcations. We employed a battery of quantitative metrics to isolate the location of such bifurcations as a function of a control parameter. Of these techniques, the Hartigans' dip statistic offered multiple advantages when used to determine the modality of a bundle's position distribution. First, the dip statistic is an inferential metric, providing an associated p-value that allowed us to estimate the bifurcation's position consistently across all the data sets. Second, the dip statistic is less sensitive to sample size and skew than other measures of a distribution's modality (29). A phenomenological bifurcation occurs when the probability distribution of a system's state, including but not limited to its position distribution, exhibits a qualitative change; for example, the distribution's modality may change (44). A third benefit of using the dip statistic is that it identified phenomenological bifurcations associated with changes in the position distribution. Although this approach might have missed modality changes in the bundle's state distribution associated with unobserved variables, such as the hair cell's transduction current, it clearly defined a boundary between an oscillatory and a quiescent regime based only on experimental observations of the bundle's motion. Although we found that noise introduced an unavoidable bias in the estimation of a deterministic bifurcation's position, we note that phenomenological bifurcations can occur at values of a control parameter that differ from those in the deterministic cases. 23 The coefficient of variation for the values of a system's state variables in a time series was previously employed to experimentally estimate the locations of Hopf bifurcations in a predator-prey system, in which the bifurcation's location was defined by the transition from a small to large value of the coefficient corresponding to the onset of spontaneous oscillations (45). By using the magnitude of the state variables rather than the time intervals between events, this coefficient of variation captures a different aspect of a dynamical system's behavior than the coefficient of variation we employ here. Because this coefficient of variation does not describe the regularity of a system's oscillations or spiking, it cannot distinguish between noisy switching in a bistable system and limit-cycle oscillations. As the mean values of the variables we study is often zero, as is the case for the Hopf normal form, it is not possible to calculate a coefficient of variation. Moreover, a change in this metric with control parameter could arise from a change in a variable's mean value rather than from a system crossing a Hopf bifurcation. Nonetheless, this metric is likely useful for bifurcation analysis of some systems especially if it were to be combined with some of the measures we utilize here. Identity of a bifurcation We developed a protocol that permits the identification of bifurcations solely on the basis of noisy time series. This diagnostic method has several appealing features. Although the approach requires data from a range of operating points near a bifurcation, it does not require external stimulation. In the study of climate change, finance, and geophysics (3, 4), among other disciplines, stimulation may be difficult or even impossible. Moreover, the method performs well at high noise levels and relies on few if any choices by the experimenter. Although the dip statistic can be employed to locate a bifurcation, it cannot be used to distinguish between types of bifurcation. To identify the bifurcation near which a system operates, we applied five additional metrics, each of which captures a different feature of the system's behavior (Fig. 6). These metrics and the analytical distributions allow us to distinguish 24 supercritical Hopf bifurcations from subcritical Hopf and SNIC bifurcations. Near a supercritical Hopf bifurcation, the RMS magnitude grows more slowly with the control parameter and the oscillation frequency and coefficient of variation are more dependent on the peak-detection threshold than for the subcritical Hopf and SNIC bifurcations. In addition, the analytical distribution often evidences a fixed point surrounded by a limit cycle when a bundle operates near a subcritical Hopf bifurcation, but never does so near a supercritical Hopf bifurcation. The similarities between simulations and observations allow us to conclude that a hair bundle possess a line of supercritical Hopf bifurcations in the high-stiffness regime The agreement between the metrics for simulations and observations also implies that a hair bundle manifests lines of either subcritical Hopf or SNIC bifurcations at low stiffnesses. Experimentally observed bundles exhibit the graded frequency response that occurs near a SNIC but not a subcritical Hopf bifurcation. However, a model bundle near a subcritical Hopf bifurcation in the presence of noise also exhibits a graded frequency response resembling that of a SNIC bifurcation (21, 22), even though no SNIC bifurcation occurs in this region of the state diagram (16). Graded frequency responses can arise from fluctuations inducing a system to cross a threshold. Moving a control parameter in a specific direction can increase the probability of crossing the threshold and consequently elevate the spiking frequency. Near a SNIC bifurcation, noise can cause a system to repeatedly cross a threshold and produce a spike. When operating in the low-stiffness regime, the model bundle instead possesses a quasi-threshold within the quiescent region near a subcritical Hopf bifurcation (Figs. S15-S18). In contrast with a true threshold that separates sub- and suprathreshold regimes, a quasi-threshold constitutes a region over which the model bundle can display spikes of all amplitudes (10). Noise can induce the model bundle's trajectory to cross this quasi-threshold, which can be very narrow, eliciting an excursion resembling an all-or-none spike. Increasing the stiffness diminishes both the amplitude of spikes and the range of constant forces over which spikes can occur. These noise-induced 25 excursions are similar to the voltage spikes or action potentials produced by neurons (6, 7, 46). For example, the Hodgkin-Huxley model possesses a quasi-threshold as does the FitzHugh- Nagumo model, a two-dimensional simplification of the Hodgkin-Huxley model that is similar in form to our qualitative hair-bundle model (16, 46, 47). In the presence of noise, the FitzHugh- Nagumo model exhibits noise-induced spikes with a frequency that depends on a control parameter. Although the graded-frequency spiking present in class 1 excitable neurons has typically been described by operation near a SNIC bifurcation (6, 7, 9), similar behavior can arise instead from model-specific dynamics near a subcritical Hopf bifurcation. Some behaviors therefore do not result solely from operation near a bifurcation, but are specific to the system in question. The presence of a graded frequency response is not sufficient to conclude that a system operates near a SNIC bifurcation. The appearance of large-amplitude oscillations as the constant force is changed is consistent with a hair bundle's crossing either a subcritical Hopf or a SNIC bifurcation. The emergence of large oscillations could alternatively arise from a third mechanism. Here a small- amplitude limit cycle is created at a supercritical Hopf bifurcation, but the amplitude of the cycle grows rapidly within an exponentially small range of control parameter values, resulting in a large-amplitude limit cycle. This phenomenon, termed a canard explosion, has been observed in a model of hair-bundle motility over a limited range of operating points (48, 49). Although canard explosions can in principle emerge in our model, it is unlikely that we observed this phenomenon in experiments. A canard explosion emerges as a sharp rise in the amplitude of oscillation with a corresponding decrease in frequency (50). In contrast, we found that large bundle oscillations appeared with a corresponding increase in frequency as the constant force declined. Although we could not reliably distinguish a noisy system operating near a SNIC bifurcation from one poised close to a subcritical Hopf bifurcation, one could in principle discriminate between these bifurcations by assessing their phase portraits. The phase portrait of a 26 system operating in the region of coexistence between a limit cycle and a fixed point near a subcritical Hopf bifurcation evidences a stable fixed point within a stable limit cycle. The phase portrait of a system near a SNIC bifurcation does not illustrate such coexistence. Using the analytic distributions as a proxy for the bundle's phase portraits, we at times found a region of increased probability within a loop as would be expected for a stable fixed point within a limit cycle (Supporting Material Section F.2). Although we require additional data to confirm that the bundle operates in a coexistence region, these results indicate that a bundle subjected to a small load stiffness can operate close to a subcritical Hopf bifurcation rather than a SNIC bifurcation. Because a single model explains the behavior of bundles for both large and small values of load stiffness as arising in part from operation near supercritical and subcritical Hopf bifurcations, respectively, it is more likely that a bundle experiences subcritical Hopf rather than SNIC bifurcations. Taken together, these data indicate a clear distinction between a bundle's operation near a supercritical Hopf bifurcation at high stiffness values and its operation close to a subcritical Hopf bifurcation at low stiffnesses. Hair-bundle function Systems operating near different bifurcations exhibit distinct behaviors. Both the proximity to and type of bifurcation can determine how a system responds to different classes of stimuli. For example, a system poised near a supercritical Hopf bifurcation responds well to periodic stimuli, whereas one operating near a SNIC bifurcation can exhibit a graded frequency response in response to changes in its control parameter. Dynamical-systems analysis therefore reveals how a system might possess different functions within different regions of its state diagram. By noting where bifurcations lie in the state diagram, one can predict how that system might function in various contexts. The stiffness of a hair bundle's load in vivo depends on the sensory organ in which it resides. For example, a mammalian auditory hair bundle tuned to 14 kHz might experience a 27 stiffness load by the tectorial membrane of over 200 mN·m-1 (43), whereas many vestibular hair bundles are coupled to otolithic membranes with a much smaller stiffness of about 1 mN·m-1 (51). A bundle's load stiffness might therefore determine its sensory role. When an auditory receptor organ imposes a high stiffness on a hair bundle, the bundle operates in the vicinity of a supercritical Hopf bifurcation. Under these conditions the bundle responds to periodic stimuli with robust amplification, sharp frequency selectivity, and a broad dynamic range (52, 53). If a bundle is instead coupled to a load of low stiffness, as might be the case in a vestibular organ, it operates in the vicinity of a subcritical Hopf bifurcation. Thermal fluctuations can induce spikes that permit the bundle to represent changes in constant force as changes in spike frequency. This graded frequency response could be useful for the detection of accelerations and gravistatic deflections. For operating points in the same region of a bundle's state diagram, a bundle can also spike in response to the beginning or end of a force step and can thus serve as a step detector owing to the quasi-threshold behavior that allows it to detect abrupt changes in force (16, 20). The dual sensory roles of individual hair bundles might be mirrored by different subpopulations of the afferent neurons that innervate them. Within vestibular organs such as the sacculus and utriculus, hair cells are innervated by afferents that discharge regularly or irregularly in the absence of stimulation (54). These neurons are classified according to their distribution of interpeak time intervals, in which regular and irregular afferents possess distributions with respectively small and large coefficients of variation. Regular afferents generate action potentials with a frequency that depends on the magnitude of a constant injected current (55), a behavior resembling that of a noisy hair bundle subjected to a low stiffness. Irregular afferents, on the other hand, respond better to periodic stimuli and display spike rates that change little with the injected current (56), similar to the behavior of an oscillating hair bundle operating in the high-stiffness regime in response to changes in constant force. Whether these neuronal subpopulations operate in the vicinity of respectively subcritical or supercritical 28 Hopf bifurcations remains to be seen. For example, as the firing rate of a regular afferent neuron decreases, its coefficient of variation correspondingly rises (55, 57). We observe the same negative correlation between the frequency and coefficient of variation as a system crosses a bifurcation indicating that the afferent neurons might also cross a bifurcation as their control parameter is changed. The algorithms presented here may permit identification of the bifurcations near which the neurons operate. Furthermore, it remains uncertain whether these two neuronal subpopulations selectively innervate hair cells with bundles operating within different functional regimes and whether afferent neurons in the auditory system possess traits similar to those of bundles operating in the high-stiffness regime. In summary, one sensory function-the detection of periodic stimuli-arises from a hair bundle's operation near a supercritical Hopf bifurcation. Two other sensory functions-the measurement of constant forces and the detection of force steps-result from a bundle's operation within the quiescent region near a subcritical Hopf bifurcation. These results highlight the remarkable flexibility of the hair bundle as a signal detector and suggest how the bundle might have evolved through changes in its operating point to serve disparate functions in various auditory, vestibular, and lateral-line organs. Acknowledgments We thank the members of our research group for comments on the manuscript. JDS is supported by grants F30DC013468 and T32GM07739 from the National Institutes of Health. AJH is an Investigator of Howard Hughes Medical Institute. 29 Authors' Contributions JDS, DÓM, and AJH designed the experiments. JDS performed the experiments and the simulations. JDS, DÓM, and AJH analyzed the data and wrote the manuscript. The authors declare no competing financial interests. References 1. 2. 3. 4. 5. 6. 7. 8. 9. Strogatz, S. 2005. Nonlinear Dynamics and Chaos. Hudspeth, A.J., F. Jülicher, and P. Martin. 2010. A critique of the critical cochlea: Hopf--a bifurcation--is better than none. Journal of Neurophysiology. 104: 1219–1229. Dijkstra, H.A. 2013. Nonlinear Climate Dynamics. Cambridge University Press. Scheffer, M., J. Bascompte, W.A. Brock, V. Brovkin, S.R. Carpenter, V. Dakos, H. Held, E.H. van Nes, M. Rietkerk, and G. Sugihara. 2009. Early-warning signals for critical transitions. Nature. 461: 53–59. Kuznetsov, Y.A. 2013. Elements of Applied Bifurcation Theory. Springer Science & Business Media. Rinzel, J., and G.B. Ermentrout. 1989. Analysis of neural excitability and oscillations In Koch, C. and Segev, I., eds., Methods in Neuronal Modeling. Izhikevich, E.M. 2000. Neural excitability, spiking and bursting. International Journal of Bifurcation and Chaos. Hodgkin, A.L. 1948. The local electric changes associated with repetitive action in a non- medullated axon. J. Physiol. (Lond.). 107: 165–181. Izhikevich, E.M. 1999. Class 1 neural excitability, conventional synapses, weakly connected networks, and mathematical foundations of pulse-coupled models. IEEE Trans Neural Netw. 10: 499–507. 10. Izhikevich, E.M. 2007. Dynamical Systems in Neuroscience. 11. Nadrowski, B., P. Martin, and F. Jülicher. 2004. Active hair-bundle motility harnesses noise to operate near an optimum of mechanosensitivity. Proc. Natl. Acad. Sci. U.S.A. 101: 12195–12200. 12. Rhode, W.S. 2007. Basilar membrane mechanics in the 6--9 kHz region of sensitive chinchilla cochleae. J. Acoust. Soc. Am. 121: 2792–2804. 30 13. Dalhoff, E., D. Turcanu, H.-P. Zenner, and A.W. Gummer. 2007. Distortion product otoacoustic emissions measured as vibration on the eardrum of human subjects. Proc. Natl. Acad. Sci. U.S.A. 104: 1546–1551. 14. de Vries, H.L. 1952. Brownian motion and the transmission of energy in the cochlea. J. Acoust. Soc. Am. 15. Hudspeth, A.J. 2014. Integrating the active process of hair cells with cochlear function. Nature Reviews Neuroscience. 15: 600–614. 16. Ó Maoiléidigh, D., E.M. Nicola, and A.J. Hudspeth. 2012. The diverse effects of mechanical loading on active hair bundles. Proc. Natl. Acad. Sci. U.S.A. 109: 1943–1948. 17. Eguíluz, V.M., M. Ospeck, Y. Choe, A.J. Hudspeth, and M.O. Magnasco. 2000. Essential nonlinearities in hearing. Phys. Rev. Lett. 84: 5232–5235. 18. Choe, Y., M.O. Magnasco, and A.J. Hudspeth. 1998. A model for amplification of hair- bundle motion by cyclical binding of Ca2+ to mechanoelectrical-transduction channels. Proc. Natl. Acad. Sci. U.S.A. 95: 15321–15326. 19. Camalet, S., T. Duke, F. Julicher, and J. Prost. 2000. Auditory sensitivity provided by self- tuned critical oscillations of hair cells. Proc. Natl. Acad. Sci. U.S.A. 97: 3183–3188. 20. 21. 22. 23. Salvi, J.D., D. Ó Maoiléidigh, B.A. Fabella, M. Tobin, and A.J. Hudspeth. 2015. Control of a hair bundle's mechanosensory function by its mechanical load. Proceedings of the National Academy of Sciences. 112: E1000–9. Fredrickson-Hemsing, L., C.E. Strimbu, Y. Roongthumskul, and D. Bozovic. 2012. Dynamics of Freely Oscillating and Coupled Hair Cell Bundles under Mechanical Deflection. Biophysical journal. 102: 1785–1792. Shlomovitz, R., L. Fredrickson-Hemsing, A. Kao, S.W.F. Meenderink, R. Bruinsma, and D. Bozovic. 2013. Low Frequency Entrainment of Oscillatory Bursts in Hair Cells. Biophysical journal. 104: 1661–1669. Salvi, J.D., D. Ó Maoiléidigh, B.A. Fabella, M. Tobin, and A.J. Hudspeth. 2015. Characterization of Active Hair-Bundle Motility by a Mechanical-Load Clamp. Mechanics of Hearing: Protein to Perception. : 030005:1–5. 24. Jacobson, M.L. 2001. Auto-threshold peak detection in physiological signals. 23rd Annual International Conference of the IEEE Engineering in Medicine and Biology Society. 25. Hartigan, J.A., and P.M. Hartigan. 1985. The dip test of unimodality. Ann. Statist. 26. Bialek, W., and H.P. Wit. 1984. Quantum limits to oscillator stability: theory and experiments on acoustic emissions from the human ear. Physics Letters A. 31 27. Longtin, A. 1990. Oscillation onset in neural delayed feedback. 28. Longtin, A., J. Milton, J. Bos, and M. Mackey. 1990. Noise and critical behavior of the pupil light reflex at oscillation onset. Phys. Rev., A. 41: 6992–7005. 29. 30. Freeman, J.B., and R. Dale. 2013. Assessing bimodality to detect the presence of a dual cognitive process. Behav Res Methods. 45: 83–97. Fraser, A.M., and H.L. Swinney. 1986. Independent coordinates for strange attractors from mutual information. Phys. Rev., A. 33: 1134–1140. 31. Cao, L. 1997. Practical method for determining the minimum embedding dimension of a scalar time series. Physica D: Nonlinear Phenomena. 110: 43–50. 32. Kennel, M., R. Brown, and H. Abarbanel. 1992. Determining embedding dimension for phase-space reconstruction using a geometrical construction. Phys. Rev., A. 45: 3403– 3411. 33. Ma, H.-G., and C.-Z. Han. 2006. Selection of Embedding Dimension and Delay Time in Phase Space Reconstruction. Front Electr Electron Eng Ch. 1: 111–114. 34. King, F.W. 2009. Hilbert Transforms:. Cambridge University Press. 35. Shera, C.A. 2003. Mammalian spontaneous otoacoustic emissions are amplitude-stabilized cochlear standing waves. J. Acoust. Soc. Am. 114: 244–262. 36. Khovanov, I.A., and L. Schimansky-Geier. 2006. Spectral analysis of noisy oscillators near hopf bifurcations. Acta Physica Polonica …. 37. Albers, D.J., and G. Hripcsak. 2012. Using time-delayed mutual information to discover and interpret temporal correlation structure in complex populations. Chaos. 22: 013111– 013111. 38. Shlomovitz, R., Y. Roongthumskul, S. Ji, D. Bozovic, and R. Bruinsma. 2014. Phase- locked spiking of inner ear hair cells and the driven noisy Adler equation. Interface Focus. 4: 20140022. 39. Howard, J., and A.J. Hudspeth. 1988. Compliance of the hair bundle associated with gating of mechanoelectrical transduction channels in the bullfrog's saccular hair cell. Neuron. 1: 189–199. 40. Martin, P., A.D. Mehta, and A.J. Hudspeth. 2000. Negative hair-bundle stiffness betrays a mechanism for mechanical amplification by the hair cell. Proc. Natl. Acad. Sci. U.S.A. 97: 12026–12031. 41. Eatock, R.A., D.P. Corey, and A.J. Hudspeth. 2003. Adaptation of Mechanoelectrical Transduction in Hair Cells of the Bullfrog's Sacculus. The Journal of Neuroscience : the 32 official journal of the Society for Neuroscience. : 1–16. 42. Le Goff, L., D. Bozovic, and A.J. Hudspeth. 2005. Adaptive shift in the domain of negative stiffness during spontaneous oscillation by hair bundles from the internal ear. Proc. Natl. Acad. Sci. U.S.A. 102: 16996–17001. 43. Ó Maoiléidigh, D., and A.J. Hudspeth. 2013. Effects of cochlear loading on the motility of active outer hair cells. Proceedings of the National Academy of Sciences. 110: 5474– 5479. 44. Arnold, L. 1998. Random dynamical systems. Berlin: Springer. 45. Fussmann, G.F., S.P. Ellner, K.W. Shertzer, and N.G. Hairston Jr. 2000. Crossing the Hopf Bifurcation in a Live Predator-Prey System. Science (New York, N.Y.). 290: 1358– 1360. 46. Fitzhugh, R. 1955. Mathematical models of threshold phenomena in the nerve membrane. Bulletin of Mathematical Biophysics. 17: 257–278. 47. Lindner, B. 2004. Effects of noise in excitable systems. Physics Reports. 392: 321–424. 48. Han, L., and A.B. Neiman. 2010. Spontaneous oscillations, signal amplification, and synchronization in a model of active hair bundle mechanics. Phys Rev E Stat Nonlin Soft Matter Phys. 81: 041913. 49. Martin, P., D. Bozovic, Y. Choe, and A.J. Hudspeth. 2003. Spontaneous Oscillation by Hair Bundles of the Bullfrog's Sacculus. Journal of Neuroscience. 23: 4533–4548. 50. Makarov, V.A., V.I. Nekorkin, and M.G. Velarde. 2001. Spiking Behavior in a Noise- Driven System Combining Oscillatory and Excitatory Properties. Phys. Rev. Lett. 86: 3431. 51. Benser, M.E., N.P. Issa, and A.J. Hudspeth. 1993. Hair-bundle stiffness dominates the elastic reactance to otolithic-membrane shear. Hearing research. 68: 243–252. 52. Martin, P., and A.J. Hudspeth. 1999. Active hair-bundle movements can amplify a hair cell's response to oscillatory mechanical stimuli. Proc. Natl. Acad. Sci. U.S.A. 96: 14306– 14311. 53. Martin, P., and A.J. Hudspeth. 2001. Compressive nonlinearity in the hair bundle's active response to mechanical stimulation. Proc. Natl. Acad. Sci. U.S.A. 98: 14386–14391. 54. Goldberg, J.M., and C. Fernández. 1971. Physiology of peripheral neurons innervating semicircular canals of the squirrel monkey. I. Resting discharge and response to constant angular accelerations. Journal of Neurophysiology. 34: 635–660. 55. Kalluri, R., J. Xue, and R.A. Eatock. 2010. Ion channels set spike timing regularity of 33 mammalian vestibular afferent neurons. Journal of Neurophysiology. 104: 2034–2051. 56. Somps, C.J., R.H. Schor, and D.L. Tomko. 1994. Vestibular Afferent Responses to Linear Accelerations in the Alert Squirrel Monkey. NASA Technical Memorandum. : 1–20. 57. Goldberg, J.M., G. Desmadryl, R.A. Baird, and C. Fernández. 1990. The vestibular nerve of the chinchilla. IV. Discharge properties of utricular afferents. Journal of Neurophysiology. 63: 781–790. Supporting References 58. Bormuth, V., J. Barral, J.-F. Joanny, F. Jülicher, and P. Martin. 2014. Transduction channels' gating can control friction on vibrating hair-cell bundles in the ear. Proc. Natl. Acad. Sci. U.S.A. 111(20): 7185–7190. 59. Freedman, D., and P. Diaconis. 1981. On the histogram as a density estimator: L 2 theory. Probability theory and related fields (Heidelberg: Springer Berlin). 57(4):453-476. 60. Ermentrout, G.B., and N. Kopell. 2006. Parabolic Bursting in an Excitable System Coupled with a Slow Oscillation. SIAM J. Appl. Math. 46(2): 233–253. 61. Ermentrout, B. 1996. Type I membranes, phase resetting curves, and synchrony. Neural Computation. 8(5): 979–1001. 62. McKennoch, S., T. Voegtlin, and L. Bushnell. 2009. Spike-Timing Error Backpropagation in Theta Neuron Networks. 21(1):9-45. 63. Izhikevich, E.M. 2007. Dynamical systems in neuroscience. The MIT Press (Cambridge, MA). 64. Treutlein, H., and K. Schulten. 1985. Noise Induced Limit Cycles of the Bonhoeffer-Van der Pol Model of Neural Pulses. Berichte der Bunsengesellschaft für physikalische Chemie. 89: 710–718. 65. Lindner, B. 2004. Effects of noise in excitable systems. Physics Reports. 392: 321–424. 66. Kralemann, B., L. Cimponeriu, M. Rosenblum, A. Pikovsky, and R. Mrowka. Uncovering interaction of coupled oscillators from data. Physical Review E, 76(055201(R)):1-4, 2007. 34 67. Kralemann, B., L. Cimponeriu, M. Rosenblum, A. Pikovsky, and R. Mrowka. Phase dynamics of coupled oscillators reconstructed from data. Physical Review E, 77(066205):1- 16, 2008. 68. Zhu, Y., Y.-H. Hsieh, R. R. Dhingra, T. E. Dick, F. J. Jacono, and R. F. Galán. Quantifying interactions between real oscillators with information theory and phase models: Application to cardiorespiratory coupling. Physical Review E, 87(2):022709, 2013. 35 Figure Legends Figure 1. The hair bundle's state diagram. (A) A theoretical state diagram specifies the behaviors of a hair bundle for various combinations of constant force (FC) and load stiffness (KL). A line of Hopf bifurcations (red) separates a region within which a hair bundle oscillates spontaneously (orange) from one in which the bundle remains quiescent (white). Bautin points (black) lie at the border between the supercritical and subcritical segments of this line. Within the spontaneously oscillatory region, a bundle's oscillations fall in amplitude and rise in frequency with an increase in load stiffness (red and blue arrows, right panels). A line of fold bifurcations (green, dashed) confines a region within which a bundle exhibits bistability (light green). The gray arrows correspond to the regions of the bundle's state diagram explored in Figures 2-5. (B) Oscillations of an experimentally observed hair bundle changed in appearance as the load stiffness increased from 300 µN·m-1 to 1000 µN·m-1 (red to purple). The position histograms and associated dip statistics shown to the right of each trace signal the presence or absence of spontaneous oscillations. The dip statistic quantifies the modality of the system's position histogram; a dip statistic of at least 0.01 with p < 10-3 implies spontaneous oscillation. (C) An experimental state diagram depicts the behavior of a hair bundle for different values of the load stiffness and constant force. White boxes correspond to operating points at which the bundle's position histogram possessed a dip statistic less than 0.01, indicating quiescent behavior. Within the territory of spontaneous oscillation, the intensities of red and blue correspond to respectively the amplitude and frequency of the bundle's oscillations calculated with a peak-detection algorithm with a threshold of 25 nm. Colored circles correspond to the traces in (B). (D) The dip- statistic values for the experimental state diagram in (C) are displayed in shades of purple. (E) Spearman's rank correlation (ρ) quantifies the relationships between amplitude (red arrow) and frequency (blue arrow) with the load stiffness (ρ(KL)) and constant force (ρ(FC)) (Table S1). For all experimental data, the stiffness and drag coefficient of the stimulus fiber were 36 respectively kSF = 109 µN·m-1 and ξSF = 142 nN·s·m-1. The proportional gain of the load clamp was 0.01. For a complete description of the bundle's experimental state diagram, see Supporting Material Section B. Figure 2. Crossing a supercritical Hopf bifurcation by adjusting the constant force. Over a range of parameter values, we compared behavior near a supercritical Hopf bifurcation with additive noise (A-G) to that of an experimentally observed hair bundle (H-N). (A) A noisy system operating near a supercritical Hopf bifurcation displayed noise-induced ringing when the control parameter was negative but oscillated spontaneously for positive values (left). When the amplitude of spontaneous oscillation surpassed that of the noise, the position histogram (middle) disclosed two peaks. When the control parameter exceeded zero, the joint probability distribution (right) for each real-valued position (X) and its Hilbert transform (XH)-the analytic distribution-was circular, with a diameter that grew with the control parameter. In all analytic distributions, red corresponds to high and blue to low probability values. Probabilities at or near zero are displayed in white. (B) The dip statistic for the position distribution is shown as function of the control parameter. A low value for the dip statistic corresponded to a unimodal distribution (green). When the control parameter exceeded zero and the amplitude of oscillation exceeded the level of the noise, the position histogram displayed two peaks (purple) corresponding to a large value of the dip statistic. The control parameter value at which the dip statistic achieved significance (shaded; p < 10-3) served as an estimate of the bifurcation's location (dashed line). (C) The RMS magnitude of the system's motion rose gradually as the control parameter increased. (D) The amplitude rose with the control parameter for peak-detection thresholds of 0.68 (red), 1.06 (yellow), and 1.45 (cyan). (E) The frequency of oscillation, with the same thresholds as in (D), grew with an increase in the control parameter until it achieved a constant value. The value of the control parameter at which the rise in frequency was steepest increased with the threshold. (F) The coefficient of variation describes the regularity of oscillation, with a 37 high value corresponding to increased variability in the interpeak time interval. The value of the control parameter at which the coefficient of variation crossed 0.5 increased with the peak- detection threshold. (G) The analytic information rose gradually with the control parameter. (H) The behavior of an experimentally observed hair bundle was transformed by changes in the constant force: as the force decreased, the amplitude of spontaneous oscillation rose (left) and the position histogram became bimodal (middle). The analytic distribution formed a loop whose diameter increased with a decrease in the constant force (right). (I) Reflecting a unimodal position histogram, the dip statistic remained small for large values of the constant force (green). When the constant force fell below 20 pN, the position histogram displayed two peaks (purple) and the dip statistic increased, defining the boundary of an oscillatory regime (shaded; p < 10-3). (J) The RMS magnitude of the bundle's motion rose gradually as the constant force decreased. (K) The amplitude increased as the constant force declined for peak-detection thresholds of 6 nm (red), 8 nm (yellow), and 10 nm (cyan). (L) The frequency of oscillation rose gradually with a decrease in constant force until it achieved a constant value near 5 pN. The force at which this rise in frequency occurred depended on the threshold value. (M) The coefficient of variation exceeded 0.5 at a value of constant force that decreased as the peak-detection threshold increased. (N) The analytic information rose gradually as the constant force fell. For all experimental data, the load stiffness was 400 µN·m-1 with a gain of 0.1. The stiffness and drag coefficient of the stimulus fiber were respectively 109 µN·m-1 and 142 nN·s·m-1. All simulations possessed noise with standard deviations of σR = σI = 0.1. The error bars represent the standard errors of the means. Figure 3. Traversing a supercritical Hopf bifurcation by altering the load stiffness. Over a range of load stiffnesses, we compared the behavior of a model hair bundle crossing a supercritical Hopf bifurcation (A-G) to that of an experimentally observed hair bundle (H-N). All of the panels depicted here display the same metrics as shown in Figure 2. (A,H) As the stiffness 38 declined, the amplitude of spontaneous oscillation rose (left), the position histogram became bimodal (middle), and the diameter of the limit cycle increased (right). (B,I) The dip statistic rose as the stiffness decreased, signalling the emergence of bimodal position histograms and the onset of spontaneous oscillations (shaded; p < 10-3). The estimated boundary of the oscillatory regime occurred on the oscillatory side of the deterministic bifurcation (dashed line). (C,J) The RMS magnitude rose gradually as the stiffness decreased. (D,K) The amplitude increased with a fall in stiffness. For these and subsequent panels, we used peak-detection thresholds of (D-F) 1 (red) and 1.5 (cyan) or of (K-M) 26 nm (red) and 34 nm (cyan). (E,L) The stiffness at which the frequency rose depended on the value of the peak-detection threshold. Spontaneous oscillations emerged with a frequency that first rose and subsequently fell with decreases in stiffness. (F,M) The stiffness at which the coefficient of variation exceeded 0.5 depended on the threshold value. (G,N) The analytic information rose gradually as the stiffness decreased. For all panels, the constant force was zero with a gain of 0.1. The stiffness and drag coefficient of the stimulus fiber were respectively 109 µN·m-1 and 142 nN·s·m-1. Dashed lines correspond to the location of the supercritical Hopf bifurcation in the absence of noise. Stochastic simulations employed noise with standard deviations of σx = σf = 0.1. The error bars represent the standard errors of the means. Figure 4. A subcritical Hopf or SNIC bifurcation at low stiffness. We compared the behavior of noisy systems near either a subcritical Hopf bifurcation (A-G) or a SNIC bifurcation (H-N) with that of an experimentally observed hair bundle subjected to changes in constant force (O-U). Each row of panels describes the same metrics as the corresponding rows in Figures 2 and 3. (A) A noisy system poised near a subcritical Hopf bifurcation displayed noise-induced bursts of high-amplitude oscillations when the control parameter fell between -0.25 and 0, a regime in which a limit cycle coexisted with a stable fixed point (left). When the control parameter exceeded zero the system exhibited high-amplitude sinusoidal oscillations. The 39 position histogram featured multiple peaks when the control parameter exceeded -0.25, a regime bounded by a saddle-node of limit cycles and a subcritical Hopf bifurcation (middle). The analytic distribution formed a cycle whose diameter changed little with the control parameter and in some cases included a stable region at the center (right). (B) The dip statistic evidenced the oscillatory boundary within the coexistence region of a subcritical Hopf bifurcation (shaded; p < 10-3). (C) The RMS magnitude grew gradually with the control parameter. (D) The amplitude rose abruptly with the control parameter for peak-detection thresholds of 1.2 (red), 1.4 (yellow), and 1.6 (cyan). (E) The frequency of oscillation rose briskly between the saddle-node of limit cycles and the subcritical Hopf bifurcation and then gradually on the exclusively oscillatory side of the Hopf bifurcation. (F) The coefficient of variation was insensitive to changes in the peak- detection threshold. (G) The analytic information rose sharply near the subcritical Hopf bifurcation. (H) A system poised near a SNIC bifurcation displayed large-amplitude spikes that increased in frequency with a rising control parameter (left). In agreement with this behavior, the position histograms became increasingly bimodal (middle). The analytic distribution disclosed a cycle whose diameter remained invariant to changes in the control parameter (right). Consistent with its spiking behavior, the system dwelled more often along one part of the cycle for low values of the control parameter. (I) The dip statistic determined the location of the deterministic SNIC bifurcation (shaded; p < 10-3). (J) The RMS magnitude grew rapidly to a constant value near the SNIC bifurcation. (K) The amplitude remained constant on the oscillatory side of the SNIC bifurcation for thresholds of 0.7 (red), 0.9 (yellow), and 1.1 (cyan). (L) The frequency of oscillation rose gradually from zero near the SNIC bifurcation for all peak-detection thresholds. (M) The coefficient of variation was insensitive to the peak-detection threshold near the SNIC bifurcation. (N) The analytic information reached a plateau for positive values of the control parameter. (O) As the constant force decreased, the spontaneous oscillations of an experimentally observed hair bundle resembled spikes of rising frequency (left), the position histograms became increasingly bimodal (middle), and the analytic distributions formed loops 40 that changed little in diameter with a locus of greater probability along one section of each cycle (right). (P) The dip statistic located the boundary of the spontaneously oscillatory region (shaded; p < 10-3). (Q) The bundle's RMS magnitude rose abruptly to a nearly constant value near the oscillatory boundary. (R) The amplitude remained nearly constant for constant forces below 25 pN for peak-detection thresholds of 20 nm (red), 27.5 nm (yellow), and 35 nm (cyan). (S) The bundle's frequency of spontaneous oscillation rose gradually from zero for constant forces below 30-35 pN. (T) The coefficient of variation remained insensitive to changes in threshold. (U) The analytic information rose sharply and then gradually as the constant force decreased. For all experimental data, the load stiffness was 50 µN·m-1 with a gain of 0.1. The stiffness and drag coefficient of the stimulus fiber were respectively 139 µN·m-1 and 239 nN·s·m-1. Black dashed lines in (B-G) and (I-N) correspond respectively to the locations of a subcritical Hopf bifurcation and a SNIC bifurcation. Pink dashed lines depict the location of a saddle-node of limit cycles bifurcation. Stochastic simulations for a subcritical Hopf bifurcation and for a SNIC bifurcation possessed noise levels of σR = σI = 0.2. The error bars represent the standard errors of the means. Figure 5. A subcritical Hopf bifurcation can be crossed by controlling the constant force. When the load stiffness remains low, a decrease in constant force advances a bundle's operating point across a subcritical Hopf bifurcation. Stochastic simulations of a model of hair-bundle motility (A-G) were compared with an experimentally observed hair bundle (H-N). The results depicted here correspond to the same metrics displayed in Figures 2-4. (A) As the constant force decreased, a model bundle exhibited noise-induced spikes of rising frequency (left), an increasingly bimodal position histogram (middle), and an analytic distribution with a cycle whose diameter changed little over a range of forces and upon which rested a locus of higher probability (right). (B) The dip statistic defined an oscillatory boundary at a control parameter smaller than the control parameters corresponding to the deterministic bifurcations (shaded; p < 10-3). (C) The model bundle's RMS magnitude rose to a nearly constant value as the constant 41 force decreased. (D) Calculated with thresholds of 1 (red) and 2 (cyan), the amplitude of the bundle's motion remained constant for forces below 1. (E) The frequency of oscillation for a model bundle rose smoothly from zero as the constant force decreased for both peak-detection thresholds. (F) The coefficient of variation remained insensitive to changes in the peak-detection threshold. (G) The analytic information rose with a decrease in constant force. (H) An experimentally observed hair bundle exhibited behaviors that accorded with those of a model bundle crossing a subcritical Hopf bifurcation in the presence of noise. As the constant force declined, the bundle displayed spikes of increasing frequency (left), a more clearly bimodal position histogram (middle), and an analytic distribution with a cycle that changed little in diameter over a range of forces and upon which rested a locus of enhanced probability (right). (I) The dip statistic defined the boundary of the oscillatory region (shaded; p < 10-3). (J) The RMS magnitude of the bundle's motion rose sharply as the constant force fell below 40-45 pN and achieved a constant value for forces below 35 pN. (K) The amplitude of the bundle's oscillations achieved a nearly constant value for forces below 35-48 pN. We employed peak- detection thresholds of 50 nm (red) and 60 nm (cyan). (L) The bundle's frequency of oscillation rose gradually from zero as the constant force decreased. (M) The constant force at which the coefficient of variation exceeded 0.5 remained relatively insensitive to changes in the peak- detection threshold. (N) The analytic information rose as the constant force decreased. For experimental data, the load stiffness was 100 µN·m-1 with a gain of 0.1. The stiffness and drag coefficient of the stimulus fiber were respectively 139 µN·m-1 and 239 nN·s·m-1. Black dashed lines correspond to the location of the subcritical Hopf bifurcation at FC = 0.66 and pink dashed lines to the location of the saddle-node of limit cycles bifurcation at FC = 0.664 in the absence of noise. Stochastic simulations of the model of hair-bundle motility possessed a stiffness of 2 and noise levels of σx = σf = 0.2. The error bars represent standard errors of the means. 42 Figure 6. Summary of metrics for the identification of bifurcations. We present schematic diagrams for each of the metrics used to identify the type of the bifurcation near which a system operates. The three columns on the left include diagrams for noisy systems crossing supercritical Hopf, subcritical Hopf, and SNIC bifurcations. The two columns on the right refer to a model hair bundle crossing either a supercritical Hopf (red) or a subcritical Hopf (blue) bifurcation. Red and blue arrows highlight the range of parameter values explored in each instance. Green arrows indicate trends in each of the statistics as a function of the control parameter. Purple arrows and purple arrowheads illustrate respectively the dependence and independence of the frequency of oscillation and coefficient of variation with changes in peak-detection threshold. For each of the panels, we highlight in color the metrics that agree with observations of bundles subjected to high- (red) and low-stiffness loads (blue). A ) C F ( e c r o f t n a t s n o C B C 80 ) 40 N p ( e c r o f t n a t s n o C 0 -40 80 40 0 ) N p ( e c r o f t n a t s n o C -40 -80 -80 400 200 1000 Load stiffness (μN·m-1) 800 600 D 0.01 Dip statistic 0.025 400 200 1000 Load stiffness (μN·m-1) 800 600 -0.5 -1 -1 -0.5 +0.5 +1 0 ρ (KL ) C F Load stiffness (KL ) KL 0.02 0.02 p(x) Amplitude (nm) 20 16 24 0.01 0.008 16 40 500 ms nm Frequency (Hz) 12 8 4 80 40 -40 -80 400 200 1000 Load stiffness (μN·m-1) 600 800 E +1 +0.5 ) C F ( ρ 0 A Supercritical Hopf bifurcation H Experimentally observed hair bundle tx I 10 nm 500 ms 1 0.8 0.6 0.4 0.2 G I A 0.9 0.7 0.5 0.3 0.1 -0.8 N I A 0.5 0.3 0.1 0.8 30 25 35 05 Constant force (pN) 15 10 20 -0.4 0 Control parameter 0.4 B c i t s i t a t s p D i 0.03 0.02 0.01 C g a m S M R D l p m A 0.7 0.5 0.3 0.1 1 0.75 0.5 .25 E F q e r F V C 0.75 0.5 0.25 J 0.03 0.02 0.01 5 4 3 2 1 c i t s i t a t s p D i ) m n ( g a m S M R K ) m n ( l p m A 7 5 3 1 L ) 9 7 5 3 1 q e r F z H ( M V C 0.75 0.5 0.25 -100.51Control parameterp(x)xxHx402780Constant force (pN)p(x)xxHx Simulated hair bundle A tx 0.1 B c i t s i t a t s p D i g a m S M R C D l p m A 0.06 0.02 0.8 0.6 0.4 0.2 1 0.75 0.5 0.25 E F q e r F 0.06 0.04 0.02 0.6 V 0.5 0.4 C G I A 1 0.8 0.6 0.4 0.2 Experimentally observed hair bundle 40 nm 200 ms 3 1 H I c i t s i t a t s p D i ) m n ( g a m S M R ) m n ( l p m A 0.07 0.05 0.03 0.01 20 15 10 5 30 20 10 J K z H L ) 4 3 2 1 q e r F ( M N 0.75 V0.5 0.25 C I A 0.7 0.4 0.1 3.5 3.25 3 Stiffness 2.75 2.5 700 900 100 Load stiffness (μN·m-1) 500 300 3.53.23.12.8Stiffnessp(x)xxHx1000533333200Load stiffness (μN·m-1)p(x)xxHx Subcritical Hopf bifurcation H A SNIC bifurcation O Experimentally observed hair bundle tx I 50 nm 3 s tx 0.05 0.03 0.01 c i t s i t a t s p D i g a m S M R J K l p m A 0.8 0.6 0.4 0.2 1 0.75 0.5 0.25 L q 0.3 0.2 0.1 e r F M N 0.75 V 0.5 0.25 C 0.2 0.1 P Q R c i t s i t a t s p D 0.03 0.02 0.01 i 40 30 20 10 ) m n ( g a m S M R ) m n ( l p m A 40 30 20 10 z H S ) 2 1.5 1 0.5 q e r F ( T U 0.75 V0.5 0.25 C 0.6 0.4 0.2 I A 3 3 B c i t s i t a t s p D i C g a m S M R D l p m A 0.05 0.04 0.03 0.02 0.01 0.9 0.7 0.5 0.3 0.1 1.25 0.75 0.5 0.25 E q e r 0.4 0.3 0.2F 0.1 F 0.75 V 0.5 0.25 C G I A 0.4 0.3 0.2 0.1 I A 0.8 -0.8 -0.4 0 0.4 Control parameter -0.8 -0.4 0 0.4 Control parameter 0.8 50 30 40 0 Constant force (pN) 20 10 -10 -0.2-0.100.4Control parameterp(x)xxHx-0.3-0.010.10.20.5Control parameterp(x)xxHx-0.123150-10Constant force (pN)p(x)xxHx27 A Simulated hair bundle H Experimentally observed hair bundle tx I 50 nm 0.15 0.1 0.05 c i t s i t a t s p D i B C g a m S M R D l p m A 1 0.5 0 1 0.5 E F q e r F V C 0.15 0.1 0.05 0 0.75 0.5 0.25 G I A 0.4 0.3 0.2 0.1 2 1 s 0.04 0.03 0.02 0.01 c i t s i t a t s p D i J ) m n ( g a m S M R K ) m n ( l p m A L ) z H ( q e r F M 24 20 16 12 8 40 30 20 10 1.6 1.2 0.8 0.4 0.75 V0.5 0.25 C I A 0.4 0.3 0.2 N 0.5 1.5 Constant force 1 0 50 30 0 40 Constant force (pN) 10 20 -10 1.81.41.30.5Constant forcep(x)xxHx50482911Constant force (pN)p(x)xxHx Supercritical Hopf bifurcation Subcritical Hopf bifurcation SNIC bifurcation e s a h P S M R l p m A q e r F V C
1901.00141
1
1901
2019-01-01T12:10:46
Transient microscopy for measuring heat transfer in single cells
[ "physics.bio-ph", "physics.app-ph" ]
Heat transfer and dissipation exists in almost any physical, chemical or biological systems. Cells, as the basic unit of life, undergo continuous heat transfer and dissipation during their metabolism. The heat transfer and dissipation within cells related to not only fundamental cellular functions and biochemical reactions, but also several important applications including heat-induced control of biological processes and treatment of diseases. Unfortunately, thus far, we still know very little about the heat transfer and dissipation properties at cellular or subcellular level. Here, we demonstrated a methodology of transient microscopy to map the heat transfer coefficients in single cells, with a temporal resolution of ~5 us and a spatial resolution of ~250 nm (close to diffraction limit). The heat transfer coefficients of different location within single cells were obtained for the first time, the inner part of cells exhibited nonuniform heat transfer properties, and it suggested a self-consistent heat transfer regulation that responds to environmental temperature.
physics.bio-ph
physics
Transient microscopy for measuring heat transfer in single cells Pei Song†1, He Gao†1, Miao Zhang1, Fan Yang1, Shan-Shan Li1, Bin Kang*1, Jing-Juan Xu*1 and Hong-Yuan Chen*1 1State Key Laboratory of Analytical Chemistry for Life Science and Collaborative Innovation Center of Chemistry for Life Sciences, School of Chemistry and Chemical Engineering, Nanjing University, 210023, China. Author Information †These authors contributed equally to this work. *Correspondence and requests for materials should be addressed to B.K. ([email protected]), J.-J.X. ([email protected]) or H.-Y.C. ([email protected]). Abstract Heat transfer and dissipation exists in almost any physical, chemical or biological systems. Cells, as the basic unit of life, undergo continuous heat transfer and dissipation during their metabolism. The heat transfer and dissipation within cells related to not only fundamental cellular functions and biochemical reactions, but also several important applications including heat-induced control of biological processes and treatment of diseases. Unfortunately, thus far, we still know very little about the 1 heat transfer and dissipation properties at cellular or subcellular level. Here, we demonstrated a methodology of transient microscopy to map the heat transfer coefficients in single cells, with a temporal resolution of ~5 µs and a spatial resolution of ~250 nm (close to diffraction limit). The heat transfer coefficients of different location within single cells were obtained for the first time, the inner part of cells exhibited nonuniform heat transfer properties, and it suggested a self-consistent heat transfer regulation that responds to environmental temperature. Introduction Heat transfer and dissipation is critical in ensuring a stable physical1, chemical2, or biological system3. Warm-blooded mammals, such as humans, consume a large part of energy for heat4, 5 and maintain body temperature via balancing heat generation and dissipation3, 6. Heat imbalance may cause an abnormal body temperature and even result in an irreversible physiological injury6, 7. Cells, as the basic unit of life where most biochemical reactions occur, undergo continuous heat generation and dissipation4, 8. The process of heat transfer and dissipation within the intracellular environment related on not only the fundamental cellular reactions,4 metabolisms5 and functions9, but also several important applications including heat-induced control of gene expression10, cancer metabolism11 and selective treatment of disease12, 13. Unfortunately, the heat transfer and dissipation properties at the cellular or subcellular level are nearly entirely unknown. Thus far, we still do not know the basic physical parameters of heat transfer within cellular environment, like heat transfer coefficient; 2 and we also do not know if cells could tune their heat dissipation capability under different environmental temperatures, just like human body does. Recently, the local temperature distribution within cells has been tentatively explored via several promising techniques14-17, with either high sensitivity of 1 -- 2 mK14 at nanoscale or high temporal resolution of several milliseconds18. Compared to measuring temperature, which is a state variable, the measurement of heat dissipation at nanoscale remains much more challenging19, especially in a single cell. The main challenge is that measuring heat dissipation requires breaking the thermal equilibrium first, creating a heat transfer process at a subcellular level, and tracking the heat dissipation dynamics before the thermal equilibrium was rebuilt. During the measurement, the disturbance to the local thermal environment should be as low as possible to avoid the effects from temperature changes. Here, we demonstrate a transient heat-transfer microscopy to map the heat transfer and dissipation properties within single cells by measuring the cooling time of a heated gold nanoparticle. The proposed method enabled visualizing heat dissipation within a single cell, with a temporal resolution of ~5 µs and a spatial resolution of ~250 nm (close to diffraction limit). The inner part of cells exhibited nonuniform heat transfer properties and a self-consistent heat transfer regulation that responds to environmental temperature. Based on these findings, we hypothesized that this cellular-level heat regulation existed in all homoeothermic animals and was an important part of body thermoregulation. 3 Results Principle of transient heat-transfer microscopy. This transient heat-transfer microscopy is based on a time-correlated pump-probe principle and a wide-field imaging configuration (Fig. 1). A 532 nm laser with 5 -- 7 ns pulses that match the maximum plasmon absorption band of gold nanoparticles, was used as a pump beam to heat the gold nanoparticle to a "hot" state, and then a ~5 µs-width pulsed white light was used as a probe beam to follow the cooling process. The sequences of pump and probe pulses were synchronized with a variable delay Δt. At time t = 0, the gold nanoparticle was rapidly heated by the pump beam to a "hot state" in several nanoseconds20-22, and a localized temperature difference ΔT was generated, thereby causing a refractive index change Δn in surrounding medium and formed a localized photothermal "nanolens"23-25 that could be optically detected. The cooling process of the "hot" gold nanoparticle could be profiled by detecting the scatting signal of the probe beam at different delays t = Δt. Based on Newton's law of cooling, the thermal time constant τ and heat transfer coefficient hm of the surrounding medium could be expressed by the following equations26, 27: Φ(t) = A∙B∙C Q ρAucAuVAu e− t τ, (1) τ = ρAucAuVAu h𝑚As , (2) where Φ(t) is the detected optical signal at time t; A, B, and C are constants; Q is the pump energy; and ρAu, cAu, VAu, and As are the density, specific heat capacity, body volume, and surface area of gold nanoparticles, respectively. 4 Figure 1. Transient heat transfer microscopy. A 532 nm pulsed laser (5 -- 7 ns, 20 Hz) was used as the pump beam to heat a gold nanoparticle (yellow sphere with red glow) within a single cell at time t = 0. The cooling process was optically monitored by using a pulsed white light as the probe beam at variable t = Δt. The heat transfer coefficient (h) of the localized cellular medium could be obtained from the cooling time (τ) according to Equation 2. Heat transfer coefficients in liquids. We measured the heat transfer coefficient in a liquid medium with single gold nanoparticles to validate the principle experiment. The detected signal Φ(t) was defined as a relative change in plasmonic scattering intensity: Φ(t) = (IΔt − Is)/Is, which originated from the refractive index change Δn caused by temperature change ΔT, where IΔt is the scattering intensity at t = Δt, and Is is the steady-state scattering intensity without a pump. Heat transfer in air was undetectable in the current system (Fig. 2a) because the Φ(t) generated in air was close to the detection limit (~±0.02) even with a strong pump. The Φ(t) was undetected in the water medium when the pump beam was switched off because no 5 "nanolens" effect was generated from surrounding nanoparticles without heating (Fig. 2b). The cooling process of a single gold nanoparticle in water and glycerol was successfully tracked when the pump and probe beams were switched on (Fig. 2c). The time profile indicated that the heat transfer coefficients for water and glycerol were 469±24 and 226±13 W∙m−2∙K−1, respectively. The influence of pump energy Q on Φ(t) was also evaluated. The initial Φ(0) at Δt = 0 exhibited a linear correlation with the pump energy (Fig. 2d), which matched well with our theory. The dependence of the measured h on the measurement conditions, including the pump energy, physical size, and surface chemistry of nanoparticles, were carefully investigated (Figs. 2e -- g). The results showed that the final h did not depend on pump energy, particle size, or surface conjugation. These results were reasonable because heat transfer coefficient h is an instinct parameter of the medium that should not depend on measurement conditions. Fig. 2e demonstrates that an increase in pump energy induced a slight increase in h, especially at relatively high pump energy of 1.2 mJ. This condition could be attributed to the local temperature change in the surrounding medium under a high pump energy because h is a temperature-dependent parameter for most liquids26. In subsequent experiments, pump energy (<0.5 mJ) and temperature change on the sample (<0.1 K) were carefully controlled to minimize thermal perturbation. Under these conditions, the localized temperature change around gold nanoparticles was <5 K, and the detection limit of our method was ~0.5 K in water and ~0.2 K in glycerol. 6 Figure 2. Single-particle-based heat transfer measurement. a,b, Curves of Φ(t) changes with time in air with pump laser on (a) and in water with pump laser off (b). c, Cooling time curves of gold nanoparticle in the surrounding medium of water (red curve) and glycerol (blue curve). d, Initial Φ(0) at t = 0 versus the pump energy. e-g, Heat transfer coefficients of water was measured under different pump energies (e), particle sizes (f), and surface conjugations (g). Heat transfer coefficients in single cells. We demonstrated that our method could be used to map heat transfer in cells by using HeLa cell line as model. The control cells without gold nanoparticles exhibited very low background signals (Φ = ~0.02 -- 0.05) closed to the detection limit (±0.02) given their low absorption cross section at 532 nm, whereas the cells with gold nanoparticles showed strong signals. In Fig. 3, images of Φ(t) at different delay times could be obtained by applying Φ(t) = (It−Is)/Is to every pixel of the dark-field (DF) image of cells with gold nanoparticles (Figs. 3a and b). The images of Φ(t) present the decay correlated with the cooling process of each 7 nanoparticle. The time profile of each pixel in the images of the Φ(t) could be exponentially fitted similar to the typical cases illustrated (Fig. 3c). The value of h at every position of the Φ(t) images could be calculated and remapped as an h map (Fig. 3d). The h map showed that cells exhibited nonuniform heat transfer properties. The heat transfer coefficient h covered a wide range from 129 W∙m−2∙K−1 to 334 W∙m−2∙K−1, with an average h = 241 W∙m−2∙K−1. Unlike in solution, the inner part of cells was a sticky and crowded environment28, 29, which contained numerous types of macromolecules and subcellular organelles30. This heterogeneous physicochemical environment might lead to the observed heterogeneity on heat transfer properties. Figure 3. Map heat transfer in single cells. a, DF image of a HeLa cell at time Δt = 5 μs. b, Φ(t) snapshot of the HeLa cell at Δt = 5 μs and Δt = 3005 μs during the cooling process. c, Cooling curves of three typical regions in the HeLa cell. d,e, Image (d) and corresponding histogram (e) of heat transfer coefficient distribution in the HeLa cell. The white dash lines in a, b, and d indicate cell contour, and the green dash lines indicate cell nuclei. 8 Heat regulation at cellular level. We explored heat transfer in cells under different environmental temperatures (Fig. 4). From a room temperature of 25 °C to a human body temperature of 37 °C, the average heat transfer coefficient of cells showed a quasi-linear increase from 231±5 W∙m−2∙K−1 to 266±6 W∙m−2∙K−1, with dh/dT = 2.9 W∙m−2∙K−2 (Fig. 4a). This phenomenon was likely a regular temperature correlation, as reflected in a cell culture medium with dh/dT = 3.3 W∙m−2∙K−2 (Fig. 4b). When the environmental temperature increased from 37 °C to 42 °C, the average heat transfer coefficient of cells remarkably increased to 313±8 W∙m−2∙K−1, with roughly estimated dh/dT = 9.5 W∙m−2∙K−2 (Fig. 4a). The heat dissipation histograms (Figs. 4c and d) clearly denote the increase in heat transfer coefficient within cells from 37 °C to 42 °C. Most cell cultures required a temperature of approximately 37 °C, and a long-term exposure to a high temperature of 42 °C would result in cell injury and even cell death31. In our measurement, the duration of high temperature was carefully controlled to several minutes to avoid irreversible cell injuries. Cell morphology remained unchanged during measurement, and consistent results could be observed when cells were cooled down to the normal temperature. Cells presented a self-consistent regulation on their heat transfer properties responsive to environmental temperature. This unique property could be conducive to maintaining the intracellular temperature within a reasonable range, although the mechanisms were unclear. The current work focused on developing fundamental methodologies and did not involve large-scale cell screening. On the basis of these results, it is reasonable to hypothesize that this cellular-level heat regulation might exist in all homoeothermic animals. 9 Further systematic verifications on a large range of cell lines are required to test this hypothesis. Figure 4. Temperature-dependent heat transfer in single cell. a, Average heat transfer coefficient of HeLa cell at environmental temperatures of 25, 30, 37, and 42 °C. b, Heat transfer coefficients of HeLa cell culture medium at environmental temperatures of 25, 30, 37, and 42 °C. c -- d, corresponding histograms of heat transfer coefficient distribution in the HeLa cell at 37 and 42 °C. Discussion The present measurement was based on single gold nanoparticles, with 40 or 80 nm in diameter. Spatial resolution mainly depends on the thickness of the thermal diffusion layer in surrounding medium. In our system, the thickness of the thermal layer was 10 related to the power of pump laser and heat diffusion time. The scale of the thermal field could be decreased to <100 nm by optimizing the pump power. The heat dissipation processes tracked in this work were typically with a thermal time constant <1 ms, and the temporal resolution (i.e., time gate width) of our transient microscopy was ~5 µs. Compared with previous techniques that measure steady-state temperatures, our strategy reached the highest temporal resolution to date. The combination of diffraction-limit imaging and transient temporal resolution capabilities in our method offered unique advantages to map the ultrafast dynamic biochemical events at cellular or subcellular level, with an nm-µs spatial-temporal resolution simultaneously. Heat dissipation within subcellular organelles could thus be explored by conjugating particles with specific targeting ligands32, 33. The combination of the proposed technique and precision temperature sensing14, 34 might further enable a comprehensive evaluation of thermogenesis and dissipation at the single-cell level. The present observations generated several unknown questions. First, extensive validations are required to verify whether this cellular-level heat regulation exists in homoeothermic and poikilothermic animals. Second, the molecular mechanism and pathways of cellular heat regulation are unknown. Fully unfolding the answers definitely requires efforts of biologists to yield relevant molecular "sensors" and "actuators" and determine the pathways in changing heat transfer. Third, heat dissipation at cell level may relate to diseases, such as cancers, and behavior under extreme conditions, such as Olympic athletes. 11 References 1. Liao, A. et al. Thermal dissipation and variability in electrical breakdown of carbon nanotube devices. Phys. Rev. B 82, 205406 (2010). 2. Meyer, J. & Reuter, K. Modeling heat dissipation at the nanoscale: An embedding approach for chemical reaction dynamics on metal surfaces. Angew. Chem. Int. Ed. 53, 4721-4724 (2014). 3. Niedermann, R. et al. Prediction of human core body temperature using non-invasive measurement methods. Int. J. Biometeorol. 58, 7-15 (2014). 4. Himms-Hagen, J. Cellular thermogenesis. Annu. Rev. Physiol. 38, 315-351 (1976). 5. Ricquier, D. Fundamental mechanisms of thermogenesis. C. R. Biol 329, 578-586 (2006). 6. Houdas, Y. & Ring, E. Human body temperature: Its measurement and regulation. (Springer Science & Business Media, 2013). 7. Bowler, K., Laudien, H. & Laudien, I. Cellular heat injury. J. Therm. Bio. 8, 426-430 (1983). 8. James, A.M. Thermal and energetic studies of cellular biological systems. (Butterworth-Heinemann, 2016). 9. Wegner, N.C., Snodgrass, O.E., Dewar, H. & Hyde, J.R. Whole-body endothermy in a mesopelagic fish, the opah, Lampris guttatus. Science 348, 786 (2015). 10. Kamei, Y. et al. Infrared laser -- mediated gene induction in targeted single cells in vivo. Nat. Methods 6, 79 (2008). 11. Vreugdenburg, T.D., Willis, C.D., Mundy, L. & Hiller, J.E. A systematic review of elastography, electrical impedance scanning, and digital infrared thermography for breast cancer screening and diagnosis. Breast Cancer Res. Tr. 137, 665-676 (2013). 12. Schroeder, A. et al. Treating metastatic cancer with nanotechnology. Nat. Rev. Cancer 12, 39 (2011). 13. Dutz, S. & Hergt, R. Magnetic nanoparticle heating and heat transfer on a microscale: Basic principles, realities and physical limitations of hyperthermia for tumour therapy. Int. J. Hyperther. 29, 790-800 (2013). 14. Kucsko, G. et al. Nanometre-scale thermometry in a living cell. Nature 500, 54-58 (2013). 15. Okabe, K. et al. Intracellular temperature mapping with a fluorescent polymeric thermometer and fluorescence lifetime imaging microscopy. Nat. Commun. 3, 705 (2012). 16. Yang, J.-M., Yang, H. & Lin, L. Quantum dot nano thermometers reveal heterogeneous local thermogenesis in living cells. ACS Nano 5, 5067-5071 (2011). 17. Nakano, M. & Nagai, T. Thermometers for monitoring cellular temperature. J. Photoch. Photobio. C 30, 2-9 (2017). 18. Chen, X. et al. Imaging the transient heat generation of individual nanostructures with a mechanoresponsive polymer. Nat. Commun. 8, 1498 (2017). 19. Hoogeboom-Pot, K.M. et al. A new regime of nanoscale thermal transport: Collective diffusion increases dissipation efficiency. Proc. Natl. Acad. Sci. U.S.A. 112, 4846-4851 (2015). 20. Qin, Z. & Bischof, J.C. Thermophysical and biological responses of gold nanoparticle laser heating. Chem. Soc. Rev. 41, 1191-1217 (2012). 21. Rashidi-Huyeh, M. & Palpant, B. Thermal response of nanocomposite materials under pulsed laser excitation. J. Appl. Phys. 96, 4475-4482 (2004). 22. Hodak, J.H., Henglein, A. & Hartland, G.V. Photophysics of nanometer sized metal particles:  Electron−phonon coupling and coherent excitation of breathing vibrational modes. J. Phys. Chem. 12 B 104, 9954-9965 (2000). 23. Chen, Z. et al. Imaging local heating and thermal diffusion of nanomaterials with plasmonic thermal microscopy. ACS Nano 9, 11574-11581 (2015). 24. Heber, A., Selmke, M. & Cichos, F. Thermal diffusivity measured using a single plasmonic nanoparticle. Phys. Chem. Chem. Phys. 17, 20868-20872 (2015). 25. Selmke, M., Braun, M. & Cichos, F. Photothermal single-particle microscopy: Detection of a nanolens. ACS Nano 6, 2741-2749 (2012). 26. Bergman, T.L. & Incropera, F.P. Fundamentals of heat and mass transfer. (John Wiley & Sons, 2011). 27. Vollmer, M. Newton's law of cooling revisited. Europ. J. Phys. 30, 1063 (2009). 28. Kuimova, M.K. et al. Imaging intracellular viscosity of a single cell during photoinduced cell death. Nat. Chem. 1, 69-73 (2009). 29. Peng, X. et al. Fluorescence ratiometry and fluorescence lifetime imaging: Using a single molecular sensor for dual mode imaging of cellular viscosity. J. Am. Chem. Soc. 133, 6626-6635 (2011). 30. Karp, G. Cell and molecular biology: Concepts and experiments 4th edition with plus set. (John Wiley & Son, 2006). 31. Lepock, J.R. Cellular effects of hyperthermia: Relevance to the minimum dose for thermal damage. Int. J. Hyperther. 19, 252-266 (2003). 32. Kang, B., Mackey, M.A. & El-Sayed, M.A. Nuclear targeting of gold nanoparticles in cancer cells induces DNA damage, causing cytokinesis arrest and apoptosis. J. Am. Chem. Soc. 132, 1517-1519 (2010). 33. Ma, X., Gong, N., Zhong, L., Sun, J. & Liang, X.-J. Future of nanotherapeutics: Targeting the cellular sub-organelles. Biomaterials 97, 10-21 (2016). 34. Toyli, D.M., de las Casas, C.F., Christle, D.J., Dobrovitski, V.V. & Awschalom, D.D. Fluorescence thermometry enhanced by the quantum coherence of single spins in diamond. Proc. Natl. Acad. Sci. U.S.A. 110, 8417-8421 (2013). Acknowledgements This work was mainly supported by the National Natural Science Foundation of China (21327902, 21535003, and 21675081), State Key Laboratory of Analytical Chemistry for Life Science (5431ZZXM1715), and Priority Academic Program Development of Jiangsu Higher Education Institutions. Author Contributions P.S. and H.G contributed equally to this work. H.-Y.C. J.-J.X. and B.K. conceived the research. 13 B.K. designed and built the transient microscope. P.S. and H.G. performed the experiments. M.Z. contributed on instrumentation operation. F.Y. and S.-S.L. contributed to sample preparation. P.S., B.K. and J.-J.X. wrote the manuscript. All authors discussed the results and commented on the manuscript. 14
1305.7030
1
1305
2013-05-30T08:09:49
One and Two-individual Movements of Fish after Chemical Exposure
[ "physics.bio-ph", "q-bio.PE" ]
Movement behavior of an indicator species, zebrafish (Danio rerio), was analyzed with one- and two-individual groups before and after treatment with a toxic chemical, formaldehyde, at a low concentration (1 ppm). After the boundary area had been determined based on experimental data, intermittency was defined as the probability distributions of the shadowing time during which data were above a pre-determined threshold and were obtained from experimental time-series data on forces and the inter-distances for one and two individuals. Overall intermittencies were similar in the boundary and central areas. However, the intermittencies were remarkably different between the one- and the two-individual groups: the single line was used to fit the data for the one-individual group whereas two phases were observed with breakpoints (approximately 10 seconds in logarithm) in the exponential fitting curves for the two-individual group. A difference in the probability distributions of shadowing time was observed "before" and "after" treatment for different areas. Intermittency patterns before and after treatment were contrasted in the center for the one-individual group whereas the difference was observed in the boundary for two-individual group. The intermittencies for the inter-distances of two individuals in the boundary and central areas were markedly different before and after treatment. When the differences between the intermittencies in the boundary and the central areas and between "before" and "after" treatment are considered, the distribution patterns of the shadowing time (scaling behaviors or intermittency patterns) should be a useful means of bio-monitoring to detect contaminants in the environment.
physics.bio-ph
physics
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 One and Two-individual Movements of Fish after Chemical Exposure Quang Kha Quach and Tae-Soo Chon* Department of Biological Sciences, Pusan National University, Busan 609-735 Hungsoo Kim Department of Biological Sciences, Pusan National University, Busan 609-735, and SPENALO National Robotics Research Center, Pusan National University, Busan 609-735 Tuyen Van Nguyen Department of Biological Sciences, Pusan National University, Busan 609-735, and Department of Mathematics, Pusan National University, Busan 609-735 Movement behavior of an indicator species, zebrafish (Danio rerio), was analyzed with one- and two- individual groups before and after treatment with a toxic chemical, formaldehyde, at a low concentration (1 ppm). After the boundary area had been determined based on experimental data, intermittency was defined as the probability distributions of the shadowing time during which data were above a pre-determined threshold and were obtained from experimental time-series data on the forces and the inter-distances for one and two individuals. Overall intermittencies were similar in the boundary and central areas. However, the intermittencies were remarkably different between the one- and the two-individual groups: the single line was used to fit the data for the one-individual group whereas two phases were observed with breakpoints (approximately 10 seconds in logarithm) in the exponential fitting curves for the two-individual group. A difference in the probability distributions of the shadowing time was observed “before” and “after” treatment for different areas. Intermittency patterns before and after treatment were contrasted in the center for the one-individual group whereas 1 24 25 26 27 28 the difference was observed in the boundary for two-individual group. The intermittencies for the inter- distances of two individuals in the boundary and the central areas were markedly different before and after treatment. When the differences between the intermittencies in the boundary and the central areas and between “before” and “after” treatment are considered, the distribution patterns of the shadowing time (scaling behaviors or intermittency patterns) should be a useful means of bio-monitoring to detect 29 contaminants in the environment. 30 31 32 PACS numbers: 02.50.Ey, 02.50.Fz, 45. 70.Cc Keywords: Intermittency, Shadowing time, Movement, Chemical stress, Force, Collective motion, 33 Boundary 34 35 36 37 38 39 40 41 42 43 44 45 46 47 *Email: [email protected]; Fax: +82-51-512-2262 I. INTRODUCTION The analysis of the response behaviors of animals has received considerable attention regarding in situ monitoring of indicator species since computational methods and interfacing techniques were introduced in the 1980’s [1-4]. Monitoring by using behavioral changes is ecologically relevant, economical and faster than monitoring by using method of chemical detection [5-7]. Due to the high degree of complexity in behavioral data, however, various computational methods have been proposed to exploring time-series data on animal movements [1, 8]: parameterization with a fractal dimension [9] and permutation entropy [10, 11], statistical methods using correlation analyses [10, 12, 13], data transform including Fourier transforms [7, 14] and wavelet analysis [15]. Considering the complexity of behavioral data, informatics has been further applied to movement patterns, including self- 2 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 organizing map [7, 14, 16] and multi-layer perception [6, 15], and is capable of identifying specific response behaviors of indicator species under chemical stress. Because of the uncertainties in behavioral patterns, the hidden Markov model has been used to analyze behavioral state changes after exposure to chemical treatment [16, 17]. However, the abovementioned reports mostly focused on data for single individuals, and not many studies were conducted on the responses of multiple individuals. Regarding group formation by multiple individuals, simulation models based on the equations of motion have been proposed to elucidate the collective behavior associated with self-propelled particle systems according to the group (i.e., overall average orientation) and the neighbor (e.g., attraction, repulsion) responses [18-23]. Group behavior models were also analyzed, and observed data were evaluated; force components of individuals in collective motion were calculated in order to explain the relationship between the individual itself, its neighbors and environmental factors [24, 25]; individual fish movements were expressed by using the mass, drag coefficient, and external forces. Recently, the importance of nearest-neighbor interactions in group formation was addressed [26, 27]. In this study, we focused on the physical forces produced by one and two individuals under stressful conditions due to chemical exposure. In order to reveal the structure property in the movement data, we addressed the probability distributions of the shadowing time in time-series force data on fish observed in a confined area. Scaling behavior has been increasingly used in analyzing movement behavioral patterns of animals in the wild and in the laboratory. Intermittency is defined as the probability distribution of the shadowing time during which the data are consecutively higher than a threshold number [28-31]. For time-series data generated from a chaotic system (e.g., attractor), intermittency exhibits a universal algebraic scaling at high frequencies with a slope approximately while it 69 exhibits an exponential scaling at lower frequencies [28, 30]. 70 71 Intermittency is among the universal mechanisms that produce chaos from a periodic orbit in a continuous way [32] and has been reported in various fields, including coordination of muscular 3 3/2 72 73 74 75 76 77 78 79 80 systems [33, 34], chemical kinetics [35, 36], laser models [37], and fluid dynamics [38]. In ecology, flow intermittency regarding biodiversity determination in stream ecosystems has been recently investigated [39, 40]. Intermittency has been further applied to behavior studies. Harnos et al. [41] analyzed scaling and intermittency in the temporal behavior of nesting gilts, reporting that the time spent by a gilt in a given form of activity had a power-law probability distribution, and showed the intermittent occurrence of certain periodic behavioral sequences to indicate a critical state. Mashnova et al. [42] investigated intermittency and a truncated power law in aphid movement and addressed the alternate appearance of fast and slow movement phases that were distinguished by a threshold value of velocity. However, intermittency in response behavior of animals under chemical stress has not been 81 extensively studied. 82 83 84 85 86 87 88 89 90 91 92 93 In addition to chemical response, we further observed individual movement at different locations in a confined area. Although test animals can move around in a more-or-less straightforward manner over a wide range, the individuals are constrained inside a confined arena within a boundary, especially for behavior monitoring within an observation arena [16]. The boundary zone was considered to be the area in which free movement would be minimally allowed, and is important for the life events, including protection and exploitation, of animals [43, 44]. We showed that the scaling behaviors of two individuals of D. rerio would be different at the boundary and the central areas of the observation arena before and after chemical treatment. Specifically, we intended to characterize intermittency in response behaviors in three different categories, 1) comparison of one and two individuals, 2) boundary and central areas, and 3) before and after chemical treatment. We analyzed the probability distributions of the shadowing time to address changes in the structure property in the movement data and found that intermittency in individual and group movement could be used as a possible means of behavioral 94 monitoring. 95 4 96 II. EXPERIMENTS 97 1. Test Organisms 98 99 One and two individuals of zebrafish, D. rerio, were observed under chemical stress. Due to vulnerability to chemical stress and availability of biological information (e.g., genomics, physiological 100 responses), the zebrafish is considered to be one of the most suitable vertebrate model organisms for 101 various biological tests [45-47], including behavior assessment [7, 16, 48]. The species has a strong 102 potential for being an indicator in risk assessment [16, 49]. Individuals of wild-type D. rerio were 103 obtained from a local fish dealer for stock population (300 individuals) and were reared for 2 weeks 104 before observation [50] at a temperature of 25 ± 1oC and pH of 7.1 ± 0.3 under a light/dark cycle of 105 14/10 h, light on at 7:00 h and off at 20:00 h [51]. Two fluorescent lights (26 J/s) were placed 50 cm 106 above the rearing container. Tap water was filtered with air stones under air compression (DT - 10F, 107 108 Chuang Xing Electric Appliances®) after dechlorination for three days. Fishes were fed dry food (Nutron Hi – Fi, PRODAC®) twice a day (once a day on weekends). Other rearing conditions are 109 described below [16]. 110 Test organisms (ages: 5 – 6 months; body lengths: 30 – 40 mm) were randomly chosen from the stock 111 population and were placed individually in a glass aquarium (300 mm × 300 mm × 300 mm ; water 112 height of 20 mm). Before observation, organisms were acclimated to the observation system for 30 113 minutes [50]. To simplify observation and minimize noise, food and oxygen were not supplied to the 114 arena during the observation period. Two 13J/s fluorescent lights were provided 50 cm above the 115 water ’s surface and the two light sources were symmetrically 32 cm away from the center over the 116 observation arena. Other rearing and observation conditions were the same as those used to rear the 117 stock population. 118 Formaldehyde (HCHO, 37wt. % solution in water, A.C.S. reagent, Gamma–Aldrich®) was used as a 119 source of stress to the test organisms. Formaldehyde is claimed to be one of most toxic environmental 5 120 hormones and a possible carcinogenic agent through bioaccumulation [52]. The chemical was directly 121 added to the water in the observation aquarium at a concentration of 1 ppm. In order to minimize noise, 122 the chemical was delivered through an injector (Pipetman® P20) connected to the observation 123 aquarium through a flexible polyethylene tube (1.85 mm in diameter and 1 m in length) after dilution 124 with a proper amount of water. 125 2. Observations and Recording 126 127 The observation system consisted of an observation aquarium, a camera (Logitech®VidTMHD), a PC (Intel® Core ™ 2 Duo CPU E4500@ 2.20GHz ), and software for tracking the motion of multiple 128 individuals. The software was developed in the Ecosystem and Behavior Lab. at Pusan National 129 University based on stereo vision [53] after evaluation with a multiple individual tracking program 130 (SynthEyes, 2008, Anderson Technologies LLC). The x-y position of each individual was continuously 131 recorded at 30 frames per second from a top view in two dimensions before (30 minutes) and after (30 132 minutes) treatment. Five-minute segments were selected for analysis according to Suzuki et al. [27] and 133 Herbert-Read et al. [54]. After treatment, fish immediately responded to olfactory stimulus from the 134 chemical for approximately 5 minutes, showing abnormal behaviors including shaking and turning. 135 Afterwards abnormal behaviors occurred less frequently. Movement tracks for the initial five minutes 136 were analyzed before and after treatment. 137 Based on preliminary research [16, 55], a time segment of 0.2 s for recording movement was selected 138 for this study. Because we aimed to observe overall movement changes of the fish specimens in two 139 dimensions in response to the chemical treatment, the 0.2 s segment was sufficiently short for 140 presenting the displacement of organism location [7, 14]. Extremely short-time response behaviors due 141 to intoxication (e.g., compulsion, trembling) may be expressed in timer shorter than 0.2 s, but this type 142 of behaviors of extremely short duration was not analyzed in this study [16]. Each movement segment 143 was determined with three points with two consecutive 0.1 s segment (0.2 s in total). The observation 6 144 was repeated 20 times for each group of one and two individuals. Mean values of the linear and the 145 angular speeds were obtained from movement segments for each individual during the observation 146 period; subsequently, the mean values were calculated from the averages of all individuals (i.e., n = 20, 147 and 40 for the one- and the two-individual groups, respectively) before and after treatment. 148 3. Computational Methods 149 3.1. Determination of boundary and central areas 150 Although test animals can move in a more-or-less straightforward manner (i.e., free run length), the 151 individuals are also located inside a confined arena for monitoring the observation arena, as stated 152 above [16]. We defined the boundary and the central areas by measuring the velocity of single 153 individuals. Figures 1(a) and (b) show the distributions of the x- and the y-component of the velocity 154 along each coordinate before treatment, respectively. In order to determine the boundary area, we 155 inspected the cumulative sum of the velocity data. In Figures 1(c) and (d), the cumulative sums of the 156 two components of the velocity from the sides of the arena are presented. 157 Subsequently, the cumulative sum were fitted with the exponential function , here A is a 158 proper amplitude and is the damping parameter, which is taken as the inverse of the width of 159 boundary. By fitting the data with the exponential function, the boundary width in x-coordinates was 160 evaluated as 19.23 mm and 10.53 mm at the left and the right sides of the x-coordinate and 20.00 mm 161 and 10.53 mm at the left and the right sides of the y-coordinate. From the evaluated value of the width 162 of boundary area from the edge of each side, the largest value, 20.00 mm, was chosen to define the 163 boundary area. The obtained value was comparable to the boundary areas empirically based on the fish 164 size [16]. 165 3.2. Real forces of each individual 7 )1(xeA 166 Based on our empirical data, we measured changes in the forces on the test individuals before and 167 after treatment. Following the framework of classical mechanics, we defined the total force on the 168 ith focal fish as the sum of two real forces, the frictional force and self-driven force : 169 170 with 171 172 (1) (2) (3) 173 where m is the mass of the fish, and µ is the friction coefficient in water. However, in our analysis, the 174 mass m is set to unity, and µ is assumed to be 0.05 [21]. 175 To calculate the self-driven force , we calculated the velocity and the acceleration of ith individual 176 at time t by using , from the movement tracks. We directly calculated the x- and the 177 y-components and the absolute force for the one-individual group while forces were calculated 178 according to center of mass, individual forces, and the relative coordinate between two individuals in 179 the two-individual group. 180 3.3. Calculation of intermittency 181 The mean value of the absolute of the force measured before treatment was used as a criterion to 182 determine the threshold for the shadowing time (Figure 2). We used one fourth of the mean value as the 183 threshold, after testing various levels of the threshold from one eighth to 2 times the mean value. One 184 fourth the mean value was most suitable in characterizing the probability distributions of the shadowing 185 time in the boundary and the central areas, as well as “before” and “after” treatment. The threshold 186 value based on the absolute value of force was also used for the x-component and the y-component. 8 iffricifdif,dfriciiifffiifma,friciifvdifiirvtiivat 187 The shadowing times and their probability distributions were expressed on a logarithmic scale. The 188 slopes and the elevations were obtained using a regression analysis [56]. The probability distribution of 189 shadowing times of long duration was also fitted to an exponential curve when breakpoints occurred in 190 intermittency [28, 30]. 191 192 Figure 3 shows the probability distributions of the shadowing time for forces on individuals observed III. RESULTS 193 in the boundary and the central areas when the time duration was selected according to the threshold 194 (95.51 mm/s2) in one- and two-individual groups. For the x- and the y-components of the forces, the 195 probability distributions of the shadowing time were overall similar between the boundary and the 196 central areas, but were different between one- and two-individual groups (Figures 3(a) – (b), (d) – (e), 197 (g) – (h), and (j) – (k)). Linearity across different shadowing times was observed for the one-individual 198 group (Figures 3(a) – (f)) whereas the linearity was not sustained and probability patterns appeared in 199 the curve for the two-individual group (Figures 3(g) – (l)). The slopes of the distribution became 200 steeper for long-time duration (i.e., right-hand side of the x-axis) for the short-time duration. For the x- 201 and the y-components, the probability distributions of the shadowing time in the boundary area 202 (Figures 3(a) – (b)) appeared to be slightly steeper than that in the central area (Figures 3(d) and (e)), 203 but no statistical difference was observed in the regression lines according to their slopes (p>0.05) [56]. 204 For the absolute forces, although the probability distributions of shadowing time were, in general, 205 similar to the x- and the y-components, a difference was observed to some degree in the shapes of the 206 probability distributions (Figures 3(c), (f), (i), and (l)). Before treatment in the boundary area, for 207 instance, the log abundance for the long-time duration appeared to spread over a broader range (i.e., 208 long foot at the right bottom corner in Figure 3(c)) whereas this type of long foot was not observed in 209 the central area. In the two-individual group, similarly, distribution patterns were different in the 210 boundary and the central areas, as well as “before” and “after” treatment (Figures 3(i) and (l)). A 9 211 detailed description of the distribution patterns, however, is beyond the scope of this study and will be 212 reported elsewhere. 213 Intermittency was further contrasted before and after treatment in Figure 4, and the probability 214 distributions of the shadowing time were fitted to lines and exponential curves because intermittency 215 exhibited a universal algebraic scaling at high frequencies and an exponential scaling at lower 216 frequencies [28, 30]. Table 1 summarizes the slopes of the lines based on regression analyses, and 217 Table 2 lists the coefficients and fittings to the exponential functions in the boundary and the central 218 areas for one- and two-individual groups. For the one-individual group, the probability distributions 219 were fitted to single lines (Figures 4(a) and (b), (d) and (e)). The slopes were similar and were in the 220 range of -1.89 – -1.91 for the x- component and -1.75 – -1.76 for the y-components before and after 221 treatment, but the difference in the slopes of the regression lines were not statistically significant 222 (p>0.05; Figures 4(a) and (b)) [56]. In the central area, however, the slopes were different for the x-and 223 the y-components of the forces. The slopes were statistically steeper for both components of the forces 224 after treatment (-1.79 – -1.80) than “before” treatment (-1.13 – -1.32) (p<0.05; Figures 4(d) and (e)), 225 indicating that the phase change in the shadowing time was more sensitive in the central area under 226 chemical stress. 227 For the absolute forces, the probability distributions of the shadowing time for one individual (Figures 228 4(c) and 4(f)) were more spread compared to those for the x-and the y-components of the forces in the 229 boundary and the central areas. The slopes appeared to be different, with statistical significance, before 230 and after treatment (Table 1). At the boundary area, slopes were steeper after treatment (-1.38) than 231 before treatment (-1.02) whereas slopes were less steep in the central area after treatment (-1.34) than 232 before treatment (-1.66) (Table 1). 233 Forces on the center of mass for the two-individual group were also calculated (Figures 4(g) – (l)). 234 Compared with the forces on the one-individual group (Figures 4(a) – (f)), the probability distributions 10 235 of the shadowing time were different, showing two phases as stated above. In both the x- and the y- 236 components, intermittency appeared to be curved with a breakpoint in the boundary (Figures 4(g) – (h)) 237 and the central (Figures 4(j) – (k)) areas whereas single lines were fitted to the intermittency curves for 238 the case of the one-individual group (Figures 4(a), (b) and (d), (e)), as stated above. The breakpoint was 239 found to be around 10 seconds, and intermittency was overall similar between the boundary and the 240 central areas for the two-individual group. It was remarkable that the difference in intermittency before 241 and after treatment was more clearly observed in the boundary area (Figures 4(g) and (h)), contrary to 242 the case of the one-individual group where the difference was only observed in the central area (Figures 243 4(d) and (e)) (Table 1). It is also noteworthy that after treatment, the elevation of the intermittency (i.e., 244 intercepts of regression lines) was lower in both the x- and the y-components in the boundary area 245 (Figures 4(g) and (h)). A statistical difference between the boundary and the central areas was observed 246 for the absolute forces (Figures 4(i) and (l), Table 1). 247 For the absolute forces on two individuals, curves were also formed in the boundary area, more 248 strongly for “after” treatment (Figure 4(i)) although higher variation was observed in probabilities 249 compared to the x and the y-components of the forces (Figures 4(g) and (h)). The breakpoint appeared 250 to slightly move toward long time duration, a little over 10 seconds (Figure 4(i)). The lines fitted to the 251 probability distributions at high frequency (i.e., before breakpoint) were statistically different before (- 252 0.89) and after (-1.19) treatment in the boundary area (Table 1). In the central area, however, single 253 lines were fitted to probability distributions across the shadowing time, and the slopes (-0.94 and -0.96) 254 were not statistically different (Table 1). We also fitted the intermittency at lower frequency (i.e., a 255 long shadowing time after breakpoint) to an exponential function [28, 30]. The coefficients were in the 256 range of 0.30 – 0.34 and exponential curves before and after treatment were not statistically different 257 when the goodness of fit between the two curves was tested according to the chi-square test [56] (Table 258 2). Although not presented in the figures, intermittency curves for velocities observed at the boundary 11 259 and the central areas were similar to the case of forces (center of mass) both “before” and “after” 260 treatment. However, the intermittency of velocities was weaker in expressing the difference between 261 “before” and “after” treatment. 262 We also calculated the relative forces between two individuals in the two-individual group (Figures 263 5(a) – (f)). Similar to force of center of mass (Figures 4(g) – (l)), two phases were observed around the 264 breakpoint of 10 seconds. The shapes of the probability distributions before and after treatment were 265 different in the boundary area while the shapes were similar in the central area. Statistical significance 266 was observed in the lines fitted to the intermittency in the x- and the y-components for the short-time 267 duration before and after treatment (Table 1) (Figures 5(a) and (b)). The slopes ranged from -0.90 to - 268 1.10 before treatment and from -1.09 to -1.63 after treatment for the x- and the y-components in the 269 boundary and the central areas. The slopes were statistically different before and after treatment in the 270 boundary area (Table 1). Although the slopes were not different in the central area, the elevations (i.e., 271 y-intercepts of the regression lines) were statistically different before and after treatment (see Ref. 56 272 for the statistical significance of the elevation in a regression line). 273 For absolute forces, the probability distributions were also different before and after treatment 274 (Figures 5(c) and (f)). A breakpoint was observed, and the point appeared to move more toward the 275 long-time duration, approximately matching 25 seconds in the boundary area. The slopes (-0.76 – -0.89) 276 of the regression lines for the absolute force were less steep compared to those of the x- and the y- 277 components (-0.90 – -1.63) in the boundary area and were comparable to those of the absolute forces (- 278 0.94 –-0.96) on the center of mass in the center area (Table 1). The coefficients (α) of the exponential 279 functions were also fitted to the probability distribution of the long shadowing time and ranged from 280 0.30 to 0. 35. The exponential curves before and after treatment were not statistically different when 281 the goodness of fit between two the curves was tested according to the chi-square test [56]. 12 282 We also checked the distribution pattern for forces for all individuals in the two-individual group 283 (Figures 5(g) – (l)). Similar to the case of intermittency of the relative force, overall probability 284 distributions were observed in two phases, single lines for short shadowing times and exponential 285 curves for long shadowing times (Figures 5(g) – (h) and (j) – (k)). According to this figure, the 286 probability distributions for the shadowing time tended to be slightly steeper in both the boundary and 287 the central areas after treatment. Difference in the slopes and the elevations were observed before and 288 after treatment for the x- and the y-components, as well as the absolute forces, and these differences 289 were statistically significant (Table 1). In the absolute forces, however, breakpoints were not clearly 290 observed in the central area (Figure 5(l)). Overall, the difference in intermittency appeared to be more 291 clearly observed in the boundary area (Table 1). Similar to the case of the relative force, intermittency 292 of individual forces was fitted to an exponential function with α values ranging from 0.30 to 0.35, and 293 the exponential curves before and after treatment were not statistically different, similar to two cases 294 above [56] (Table 2). 295 We also calculated the probability distributions of the shadowing time for two individuals’ inter- 296 distance. The difference was outstanding in the boundary area before and after treatment ; the curve 297 became rapidly steeper after the breakpoint (Figure 6(a)). Similar to the case of forces, the break point 298 was formed around 10 seconds. In the center, however, single lines were fitted both “before” and “after” 299 treatment. The slopes of the intermittency appeared to be flat, ranging from 0.22 to -0.26 in the 300 boundary area. The slope after treatment, however, became steeper (-0.80) than the slope before 301 treatment (-0.47) in the central area (Figure 6(b)). The slopes before and after treatment were 302 statistically different for both the boundary and the central areas (Table 1, Figure 6). Exponential 303 functions were fitted to the intermittency after treatment, with α = 0.26 (R2=0.71) in the boundary area, 304 according to chi-square-test goodness of fit [56] (Table 2). 305 IV. DISCUSSIONS AND CONCLUSIONS 13 306 It was remarkable that the data structure was fundamentally different between single and two 307 individuals. The breakpoints with two phases in intermittency were observed for short and long 308 shadowing times in the two-individual group (Figures 3(g), (h), (j) and (k)) whereas single lines were 309 presented in the one-individual group (Figures 3(a), (b), (d), and (e)). The linearity and the breakpoints 310 were consistently observed both “before” and “after” treatment (Figures 3 – 5). This indicates that 311 pairwise interaction between two individuals played a key role in determining movement data structure. 312 Recently, the importance of the nearest-neighbor relationship in group behavior was reported. Herbert- 313 Read et al. [54] demonstrated the importance of repulsion and response to a single nearest neighbor in 314 fish group–behavior dynamics. Pairwise interactions are important in qualitatively capturing the correct 315 spatial interactions in small groups of fish when compared with the observed data [57]. Our study 316 indirectly supports the significance of two-individual interactions in group formation. 317 In addition, the intermittency patterns were substantially different “before” and “after” chemical 318 exposure for different areas in the observation arena (Figures 4 – 6). The probability distributions of the 319 shadowing time were different before and after treatment in the center area for the one-individual group 320 whereas the difference was observed in the boundary area for the two-individual group. Regarding 321 behavioral-state changes (i.e., transition probability of different movement patterns), no qualitative 322 difference was observed between the boundary and the central areas [16]. Indeed, the overall patterns 323 of intermittency were similar in the boundary and the central areas (Figure 3). However, response to 324 chemical stress appeared differently according to the organism’s location in the arena. Especially, the 325 inter-distances between two individuals were markedly different “before” and “after” treatment (Figure 326 6). This further indicates that pairwise interactions are strongly reflected in the spatial dynamics in the 327 boundary area, suggesting emergence of new property in the movement data structure in responding to 328 neighbors nearby edge areas under stressful conditions. To the best of the authors’ knowledge, this is 329 the first report observing differences in the intermittency of forces on individuals in the boundary and 14 330 the central areas. Further study, however, is required in both the computational and the biological 331 aspects. 332 The existence of breakpoints in two-individual groups (Figures 3 – 5) also reflects critical time 333 duration for characterizing collective motion. Considering that the slopes for intermittency at short 334 times were near -1.5 and the slopes became steeper, the time duration of 10 seconds may be due to 335 behaviors stemming from the association of two individuals in a confined area (e.g., approach, 336 communication). The breakpoints moved toward longer time duration in the case of absolute forces 337 (Figures 5(c) and (i)). Currently, the mechanism of breakpoint formation is not known. This time 338 duration may also be due to an output from physiological networks [58]. In biological aspects, 339 physiological and/or molecular genetics networks could be investigated; how stereotypic changes in 340 behavioral patterns could originate from integrative actions of neural and endocrine systems [50]. 341 However, the detailed mechanism is currently unknown and more research may be required in this 342 direction in the future. 343 Considering the difference in the intermittency patterns at different locations before and after 344 treatment, especially in the boundary area, the probability distributions of the shadowing time could be 345 utilized as a useful means of monitoring chemical stress. Intermittency in the inter-distance between 346 two individuals was remarkably different between “before (i.e., strong curves with a breakpoint)” and 347 “after (i.e., single line)” treatment in the boundary, as shown in Figure 5(a). 348 In this study, we did not use the abundance data for the minimal time duration (i.e., the first 349 probability matching to the shortest shadowing time); the points were not maximal in all cases (Figures 350 3 – 6). For instance, the point matching minimum time duration in Figure 4(a) showed abundance less 351 than the abundance shown by the second shortest shadowing time. Considering the negative value (- 352 1.5) of the intermittency [30], the abundance should be theoretically maximized at the shortest 353 shadowing time. The somewhat lower abundance at the minimum shadowing time indicates that the 15 354 shortest shadowing time is expressed in a reserved manner biologically, and this may stem from the 355 physiological and behavioral nature of the organisms. However, the reason is currently unknown. In 356 some cases, sufficient data to evaluate intermittency were not recorded. For instance, the intermittency 357 applied to the relative force in the central area, insufficient data points were collected in the central area 358 (Figures 5(d) and (e)). This may be due to the fact that more data points were recorded in the boundary 359 area. Considering that an acute response due to the olfactory stimulus of formaldehyde were generally 360 observed within 5 minutes as stated above, the observation time may not be extended due to weaker 361 response behaviors after 5 minutes, but the replication number may be increased. More data need to be 362 accumulated in a future study. 363 In this study, only one concentration of the chemical was tested. More research is needed at different 364 concentrations of chemicals in order to determine the fish’s behavioral response to an increase in stress 365 levels. In the future, more than two individuals could be tested, and the contributions of additional 366 neighbors to group formation could be more closely investigated. 367 In conclusion, the intermittency of forces and inter-distances in one- and two-individual groups 368 effectively addressed the structural changes in collective motion. Whereas linearity was observed in the 369 probability distributions of the shadowing time for the one-individual group, two phases with 370 breakpoints were measured for two-individual group consisting linearity (the short shadowing time) 371 and exponential function (the long shadowing time). Furthermore, the effect of chemical stress was 372 demonstrated by using difference between the intermittencies in the boundary and the central areas. 373 Differences in the intermittency patterns appeared more clearly in the center for the one-individual 374 group, but the differences were more effectively presented in the boundary for the two-individual 375 group. Changes in the probability distributions of the shadowing time suggested that the pairwise 376 association between two individuals is essential in collective motion and group formation. The 377 sensitivities in the intermittencies evaluated for the one- and the two-individual groups in response to 16 378 toxic chemicals can be utilized as a means of behavioral monitoring to detect contaminants in the 379 environment. 380 ACKNOWLEDGMENT 381 This research was supported by the Basic Science Research Program through the National Research 382 Foundation of Korea (NRF) funded by the Ministry of Education, Science and Technology (2011- 383 0012960). 384 REFERENCES 385 386 387 388 389 390 391 [1] A. D. Lemly and R. J. F. Smith, Ecotox. and Environ. Safe. 11, 211 (1986). [2] H. Dutta, J. Marcelino, and C. Richmonds, Arch. Int. Phys. 100, 331 (1992). [3] T. S. Chon, N. Chung, I. S. Kwak, J. S. Kim, S. C. Koh, S. K. Lee, J. B. Leem, and E. Y. Cha, Environ. Monit. Assess. 101, 2 (2005). [4] J. A. Macedo-Sousa, J. L. T. Pestana, A. Gerhardt, A. J. A. Nogueira, and A. M. V. M. Soares, Chemosphere 67, 1665 (2007). [5] A. Gerhardt, Biomonitoring of polluted water: reviews on actual topics (Trans Tech Publications, Uetikon- 392 Zuerich, 1999). 393 394 395 396 397 [6] I. S. Kwak, T. S. Chon, H. M. Kang, N. I. Chung, J. S. Kim, S. C. Koh, S. K. Lee, and Y. S. Kim, Environ. Pollut. 120, 672 (2002). [7] Y. S. Park, N. I. Chung, K. H. Choi, E. Y. Cha, S. K. Lee, and T. S. Chon, Aquat. Toxicol. 71, 217 (2005). [8] B. J. Lawrence and R. J. F. Smith, J. Chem. Ecol. 15, 210 (1989). [9] C. W. Ji, S. H. Lee, K. H. Choi, I. S. Kwak, S. G. Lee, E. Y. Cha, S. K. Lee, and T. S. Chon, Int. J. Ecodyn. 398 2, 28 (2007). 399 [10] Y. Liu, T. S. Chon, H. K. Baek, Y. H. Do, J. H. Choi, and Y. D. Chung, Mod. Phys. Lett. B 25, 1135 400 (2011a). 401 402 403 [11] Y. Li, J. M. Lee, T. S. Chon, Y. Liu, H. Kim, M. J. Bae, and Y. S. Park, Mod. Phys. Lett. B 27, 135 (2013). [12] P. C. Tobin and O. N. Bjørnstad, J. Ani. Ecol. 72, 461 (2003). [13] S. Dray, M. Royer-Carenzi, and C. Calenge, Ecol. Res. 25, 675 (2010). 17 404 [14] T. S. Chon, Y. S. Park, K. Y. Park, S. Y. Choi, K. T. Kim, and E. C. Cho, Appl. Entomol. and Zool. 39, 79 405 (2004). 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 [15] C. K. Kim, I. S. Kwak, E. Y. Cha, and T. S. Chon, Ecol. Model. 195, 62 (2006). [16] Y. Liu, S. H. Lee, and T. S. Chon, Ecol. Model. 222, 2192 (2011b). [17] T. V. Nguyen, Y. Liu, I. H. Jung, T. S. Chon, and S. H. Lee, Mod. Phys. Lett. B 25, 1144 (2011). [18] T. Vicsek, A. Czirók, E. Ben-Jacob, I. Cohen, and O. Shochet, Phys. Rev. Lett. 75, 1227 (1995). [19] S. V. Viscido, M. Miller, and D. S. Wethey, J. Theor. Biol. 208, 315 (2001). [20] J. K. Parrish, S. V. Viscido, and D. Grünbaum, Biol. Bull. 202, 296 (2002). [21] S. H. Lee, H. K. Pak, and T. S. Chon, J. Theor. Biol. 240, 250 (2006). [22] A. S. Kane, J. D. Salierno, G. T. Gipson, T. C. A. Molteno, and C. Hunter, Water Res. 38, 3993 (2004). [23] S. Y. Ha, S. Jung, and M. Slemrod, J. Differ. Equations 252, 2563 (2012). [24] N. S. Matuda, Bull. of the Jpn. Soc. Sci. Fisher. 46, 689 (1980). [25] S. H. Lee, J. of Asia-Pac. Entomol. 16, 12(2013). [26] K. N. Tsutomu, T. Takagi, K. Yamamoto, and T. Hiraishi, Nippon Suisan Gakkaishi 59, 1280 (1993). [27] K. Suzuki, T. Takagi, and T. Hiraishi, Fish. Res. 60, 4 (2003). [28] J. E. Hirsch, B. A. Huberman, and D. J. Scalapino, Phys. Rev. A 25, 519 (1982). [29] W. C. B. David, K. Sauer, and R. J. Otis, J. Environ. Eng. Div. 102, 790 (1976). [30] Y. Do, Y. C. Lai, and Z. Liu, E. J. Kostelich, Phys. Rev. E 67, 202 (2003) . [31] Y. Do and Y. C. Lai, Europhys. Lett. 67, 915 (2004). [32] P. Manneville and Y. Pomeau, Phys. Lett. A 75, 1 (1979). [33] N. F. S. Lawrence, Biophys. J. 8, 252 (1968). [34] P. Gawthrop, I. Loram, M. Lakie, and H. Gollee, Biol. Cybern. 104, 32 (2011). [35] Y. Pomeau, J. C. Roux, A. Rossi, S. Bachelart, and C. Vidal, J. Physique Lett. 42, 272 (1981). [36] I. M. De la Fuente, L. Martinez, and J. Veguillas, Biosystems 39, 88 (1996). [37] G. J. de Valcárcel, E. Roldán, V. Espinosa, and R. Vilaseca, Phys. Lett. A 206, 360 (1995). [38] Y. Pomeau and P. Manneville, Commun. in Math. Phys. 74, 190 (1980). [39] T. Datry, D. Arscott, and S. Sabater, Aquat. Sci. 73, 454 (2011). 18 431 432 433 434 [40] M. T. Bogan, K. S. Boersma, and D. A. Lytle, Freshwater Biol. 10, 1111 (2013). [41] A. Harnos, G. Horváth, A. B. Lawrence, and G. Vattay, Physica A 286, 312 (2000). [42] A. Mashanova, T. H. Oliver, and V. A. A. Jansen, J. R. Soc. Interface 7, 199 (2010). [43] R. Jeanson, S. Blanco, R. Fournier, J. L. Deneubourg, V. Fourcassié, and G. Theraulaz, J. Theor. Biol. 225, 435 443 (2003). 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 [44] M. Martin, F. Bastardie, D. Richard, F. Burel, and C. R. Acad. Bulg. Sci. 324, 1029 (2001). [45] J. R. Fetcho and K. S. Liu, Ann. NY Acad. Sci. 860, 334 (1998). [46] G. R. Blaser R, Behav. Res. Methods 38, 456 (2006). [47] E. Levin, Z. Bencan, and D. Cerutti, Physiol. and Behav. 90, 55 (2007). [48] S. Kato, J. Neurosci. Methods 134, 2 (2004). [49] H. A. Swain, C. Sigstad, and F. M. Scalzo, Neurotoxicol. and Teratol. 26, 726 (2004). [50] R. Gerlai, V. Lee, and R. Blaser, Pharmacol. Biochem. Behav. 85, 753 (2006). [51] N. Miller and R. Gerlai, Behav. Brain Res. 184, 157 (2007). [52] N. Noordiana and Y. C. B. Farhana, Int. Food Res. J. 18, 125 (2011). [53] C. Xia, Y. Li, T. S. Chon, and J. M. Lee, Indust. Elec. ISIE, 909 (2009). [54] J. E. Herbert-Read, A. Perna, R. P. Mann, T. M. Schaerf, D. J. T. Sumpter, and A. J. W. Ward, P. Natl. A. Sci. 108, 18726 (2011). [55] R. D. Collins, R. N. Gargesh, A. D. Maltby, R. J. Roggero, M. K. Tourtellot, and W. J. Bell, Physiol. Entomol. 19, 166 (1994). [56] J. H. Zar, Biostatistical analysis (Prentice hall, Upper Saddle River, New Jersey, 1999). [57] Y. Katz, K. Tunstrøm, C. C. Ioannou, C. Huepe, and I. D. Couzin, P. Natl. A. Sci. 108, 18720 (2011) . [58] I. D. Couzin, J. Krause, R. James, G. D. Ruxton, and N. R. Franks, J. Theor. Biol. 218, 2 (2002). 19 453 Table 1 Estimates of the slopes and elevations by applying a regression analysis to the intermittency 454 of forces for movement of zebrafish in one- and two-individual groups in the boundary and central 455 areas before and after chemical treatment. No. of indi Type Treat X Boundary Y Absolute One Indi Before -1.89±0.20 -1.76±0.19 -1.02±0.27 After -1.91±0.19 -1.75±0.19 -1.38±0.26* Center of Before -1.11±0.17 -1.10±0.11 -0.89±0.11 mass After -1.61±0.17* -1.48 ±0.14* -1.19±0.16* Before -1.09 ±0.13 -1.1±0.09 -0.79±0.18 Relative Two After -1.63±0.16* -1.41±0.09* -0.89±0.14① Before -1.35±0.11 -1.28±0.08 -0.71±0.15 Indi After -1.61±0.17* -1.52±0.09* -0.89±0.16* X Center Y Absolute -1.32±0.16 -1.13±0.16 -1.66±0.49 -1.79±0.11* -1.80±0.18* -1.34±0.11* -1.08±0.15 -1.17±0.18 -0.94±0.26 -1.26±0.32* -1.14±0.26* -0.96±0.31 -0.90±0.19 -1.07±0.15 -0.76±0.32 -1.30±0.22② -1.09±0.20③ -0.87±0.40* -0.94±0.10 -1.07±0.12 -0.79±0.06 -1.03±0.19④ -1.02±0.17⑤ -1.03±0.07* 456 457 458 459 * Indicates statistical significance “before” and “after” treatment based on the different slopes of the regression lines (p<0.05) [56]. Numbers in circles present statistical significances “before” and “after” treatment based on the different elevations in the regression lines (p<0.05) [56] ①-2.07/-1.50, ②-1.15/-0.97, ③-1.08/-0.81, ④-1.40/-1.17 and ⑤-1.47/-1.10, before /after treatment, respectively. 20 460 461 462 463 464 465 466 Table 2 Estimates of the coefficient (α) of the exponential function and the goodness of fit (chi-square test) applied to the intermittency of forces of zebrafish in one- and two-individual groups before and after treatment. Type Treat Test Boundary X Y Center of mass Before † 0.31 (0.42) 0.30 (0.41) After † 0.35 (0.61) 0.34 (0.62) χ2 †† 10.66 (0.15) 10.14 (0.18) Before 0.30 (0.64) 0.31 (0.45) Relative After 0.35 (0.79) 0.33 (0.85) χ2 10.28 (0.17) 11.72 (0.11) Before 0.33 (0.48) 0.33 (0.52) Indi After 0.35 (0.58) 0.35 (0.52) χ2 3.79 (0.80) 3.19 (0.87) Center X Y 0.30 (0.59) 0.30 (0.36) 0.30 (0.30) 0.30 (0.25) 0.63 (1.00) 1.79 (0.97) 0.30 (0.56) 0.31 (0.27) 0.31 (0.30) 0.31 (0.56) 5.61 (0.59) 11.89 (0.10) 0.30 (0.34) 0.30 (0.29) 0.30 (0.61) 0.31 (0.41) 5.30 (0.63) 7.67 (0.47) † Numbers in parentheses indicate the R2 value according to the coefficient estimate of the functions (exponential decay [56]). †† Numbers in parentheses present the probability according to chi-square test’s goodness of fit between two exponential functions, one before and one after treatment [56]. 21 467 468 Fig. 1. Velocity distribution and its cumulative sum along each axis of fish movement in the 469 observation arena in defining the boundary area: (a) x-component, (b) y-component, (c) cumulative 470 sum of the x-component, and (d) cumulative sum of the y-component. Solid curves at the boundary (c) 471 and (d) indicate the exponential curves fitting the data from 0 mm to 50 mm. 22 472 473 Fig. 2. Time series of the absolute value of force for one individual before treatment in the center for 474 various shadowing times (t) and a threshold (dashed line) to determine the shadowing time. 23 475 476 Fig. 3. Intermittency patterns for forces on one individual before (a, b, and c) and after (d, e, and f) 477 treatment, and those for two individuals before (g, h, and i) and after (j, k, and l) treatment in the 478 boundary (blank squares) and the central (red circles) areas. The probability distributions of the 479 shadowing time were fitted to single lines for the one-individual group whereas they were matched to 480 exponential curves for the two-individual group. 481 24 482 483 Fig. 4. Intermittency patterns for the forces on the one-individual group in the boundary area ((a) x- 484 components, (b) y-component, and (c) absolute value) and in the central area ((d) x-component, (e) y- 485 component, and (f) absolute value), and those for force at the center of mass of the two-individual 486 group in the boundary area ((g) x-component, (h) y-component, and (i) absolute value) and the central 487 area ((j) x-component, (k) y-component, and (l) absolute value). Intermittency patterns before and after 488 treatment were different in the center for the one-individual group whereas the difference was observed 489 in the boundary for the two-individual group. Solid and dotted lines fitting “before” and “after” 490 treatment, respectively. 25 491 492 Fig. 5. Intermittency patterns for the forces before (blank squares) and after (red circles) treatment in 493 the two- individual group. The relative force on individuals in the boundary area ((a) x-component, (b) 494 y-component, and (c) absolute value) and in the central area ((d) x-component, (e) y-component, and (f) 495 absolute value), and those on two individuals in the boundary area ((g) x-component, (h) y-component, 496 and (i) absolute value) and in the central area ((j) x-component, (k) y-component, and (l) absolute 497 value). Differences in the intermittency patterns before and after treatment were more clearly observed 498 in the boundary for relative forces whereas the difference was equally observed in the boundary and the 499 center for individual forces. Solid and dotted lines fitting “before” and “after” treatment, respectively. 26 500 501 Fig. 6. Intermittency patterns for the inter-distance between two individuals before (blank squares) 502 and after (red circles) treatment in (a) the boundary (slopes before (-0.22 ± 0.08) and after (-0.26 ± 0.12) 503 treatment), and (b) the center (slopes before (-0.47 ± 0.06) and after (-0.80 ± 0.11) treatment). The 504 intermittency pattern was markedly different after treatment in the boundary with a breakpoint clearly 505 separating flat (for short shadowing time) and steep (for long shadowing time) slopes . Solid and dotted 506 lines fitting “before” and “after” treatment, respectively. 27
1509.06454
1
1509
2015-09-22T03:18:19
Stabilization of helical macromolecular phases by confined bending
[ "physics.bio-ph", "cond-mat.soft" ]
By means of extensive replica-exchange simulations of generic coarse-grained models for helical polymers, we systematically investigate the structural transitions into all possible helical phases for flexible and semiflexible elastic polymers with self-interaction under the influence of torsion barriers. The competing interactions lead to a variety of conformational phases including disordered helical arrangements, single helices, and ordered, tertiary helix bundles. Most remarkably, we find that a bending restraint entails a clear separation and stabilization of the helical phases. This aids in understanding why semiflexible polymers such as double-stranded DNA tend to form pronounced helical structures and proteins often exhibit an abundance of helical structures, such as helix bundles, within their tertiary structure.
physics.bio-ph
physics
Stabilization of helical macromolecular phases by confined bending Matthew J. Williams1, ∗ and Michael Bachmann1, 2, 3, † 1Soft Matter Systems Research Group, Center for Simulational Physics, The University of Georgia, Athens, GA 30602, USA 2Instituto de F´ısica, Universidade Federal de Mato Grosso, Cuiab´a (MT), Brazil 3Departamento de F´ısica, Universidade Federal de Minas Gerais, Belo Horizonte (MG), Brazil (Dated: September 26, 2018) By means of extensive replica-exchange simulations of generic coarse-grained models for helical polymers, we systematically investigate the structural transitions into all possible helical phases for flexible and semiflexible elastic polymers with self-interaction under the influence of torsion barriers. The competing interactions lead to a variety of conformational phases including disordered helical arrangements, single helices, and ordered, tertiary helix bundles. Most remarkably, we find that a bending restraint entails a clear separation and stabilization of the helical phases. This aids in understanding why semiflexible polymers such as double-stranded DNA tend to form pronounced helical structures and proteins often exhibit an abundance of helical structures, such as helix bundles, within their tertiary structure. PACS numbers: 64.70.-p, 82.35.Lr, 83.10.Tv, 87.15.Cc Helical segments are ubiquitous secondary structures occurring in most macromolecular systems. The forma- tion of helical structures is typically attributed to the formation of hydrogen bonds along the backbone of lin- ear polymers, but it is also known that helices are among the few generic geometries that a linelike topology can form if an ordering principle (such as a many-body con- straint) is present [1 -- 3]. In seminal works, Zimm and Bragg (ZB) [4, 5] showed that the crossover between disordered random coil struc- tures and ordered helical conformations can be described by a one-dimensional Ising-like model. Therefore, while short-range cooperativity can lead to structural ordering, in the ZB model this process is not a phase transition in the strict thermodynamic sense [6, 7]. However, since biologically relevant macromolecules are finite systems (on an effectively mesoscopic scale), the thermodynamic interpretation of structural transitions in such systems must address finiteness effects accordingly [8]. Primary effects of cooperativity can be addressed by generic effective-potential models that allow for the qual- itative description of helix -- coil transitions [9 -- 12]. Dom- inant nonbonded interactions support the formation of tertiary structures including single helices, helix bundles, collapsed globules, or random coils [13 -- 20]. It has been shown recently that the alignment of secondary struc- tures in a tertiary protein fold can be understood as a simple two-state process [21]. For a generic flexible polymer chain, in a crystalliza- tion process succeeding the chain collapse, ordered struc- tures emerge that are substantially different from ter- tiary structures known from realistic biomolecules and typically do not possess secondary structures [22, 23]. If bending restraints and nonbonded interaction compete with each other, as it is the case in self-interacting semi- flexible polymers, the ordered structures are known to be rodlike bundles or toroids [24]. However, less is known about the influence of effec- tive bending restraints upon transition pathways toward helical structures. A systematic analysis of the forma- tion and separation of helical phases in phase space in the presence or absence of bending restraints has not yet been performed. In this study, we investigate the rel- evance of this restraint for the separation of structural phases by means of replica-exchange Monte Carlo com- puter simulations for coarse-grained flexible and semiflex- ible polymer models. By scanning the spaces of torsion parameter strength and temperature, we construct the hyperphase diagrams for entire classes of helical macro- molecules, which allows us to distinguish the different pathways to the helical folds and enables us to judge the significance of bending restraints in biomacromolecules. We employ a generic coarse-grained bead-spring model for elastic, self-interacting polymers with torsional inter- action. The polymer is represented by a linear chain of N monomers. The bending energy of flexible poly- mers is zero. For semiflexible polymers, excitations of the bond angle formed by successive bonds are subject to an energetic penalty. Torsion is induced by an out- of-plane torsion angle between three successive bonds. Nonbonded monomers interact via long-range attractive and short-range repulsive van der Waals forces, mod- eled by the Lennard-Jones (LJ) potential. The energy of a conformation X = {x1, x2, . . . , xN}, where xi is the location of the ith monomer, is given in units of the LJ energy scale ǫ by E(X)/ǫ = Pi>j+1 vLJ(rij ) + sr Pi vbond(ri i+1)+sθ Pk vbend(θk)+sτ Pl vtor(τl). The dimensionless Lennard-Jones potential with cut-off is given by vLJ(r) = 4[(σ/r)12 − (σ/r)6] − vc if r < rc = 2.5σ, where r is the distance between two nonbonded monomers and vc ≈ −0.0163 is a constant shift to en- sure vLJ(rc) = 0. For r > rc, we set vLJ(r) = 0. 2 values K = (98/5)ǫr2 0 and R = (3/7)r0. The pay-off for bending the chain is vbend(θ) = 1−cos(θ−θ0), where θ0 is the bond angle in the ground state. The bending energy scales with sθ = Sθ/ǫ. For the simulations of the flexible polymer Sθ = 0 (no bending restraint), whereas for the semiflexible polymer Sθ = 200ǫ was chosen. Eventually, the torsion potential is vtor(τ ) = 1 − cos(τ − τ0), with the dihedral torsion angle τ and its equilibrium value τ0. The relative energy scale is sτ = Sτ /ǫ. The choice of reference angles τ0 = 0.873 and θ0 = 1.742 allows for helical segments in the ground-state structures that re- semble right-handed α helices with about 4 monomers per turn. In the following, ǫ, r0, and kB are set to unity. For the simulation of polymers with up to 60 monomers, replica-exchange Monte Carlo parallel tempering simula- tions were performed [26 -- 29]. The propensities of polymers with (Sθ > 0) and with- out bending restraint (Sθ = 0) to form stable helical structures are investigated under thermal conditions con- trolled by the canonical heat-bath temperature T . The variation of the torsion strength Sτ enables the study of an entire class of helical polymers. Representations of transition channels in generalized ensembles have turned out to be beneficial [8, 30 -- 32]. Therefore, we discuss the folding channels of the helical polymers in the mul- tiplicative canonical ensemble provided by the parallel tempering method. We introduce a pair of order param- eters which are effectively defined by the average total energies per monomer of the nonbonded LJ interactions between all monomers and their neighbors up to 6 bonds away, q1(X) = ǫ 1 N N −2 X i=1 N X j=i+2 Θ6,j−i vLJ(rij ), (1) and all others, q2(X) = ǫ 1 N N −2 X i=1 N X j=i+2 Θj−i,7 vLJ(rij ), (2) where Θkl = 1 if k ≥ l and zero otherwise. In a single long helix, q1 is minimal and q2 maximal, whereas for helix bundles with increasing number of segments q2 gets smaller and q1 larger. For a 40mer, Fig. 1 depicts for a selection of tor- sion strength values Sτ the distributions of conformations found in the generalized ensemble that covers the tem- perature interval T ∈ [0.1, 2.0]ǫ/kB in (q1, q2) space. The left column of figures [(a)-(d)] shows the helical tran- sition pathways for the bending-restrained semiflexible polymer (Sθ = 200ǫ) and the right column [(e)-(h)] for the unrestrained flexible polymer (Sθ = 0). The lightgray region represents the area in (q1, q2) space, in which con- formations were found at all temperatures and torsion strengths in the simulations. This distribution, which is independent of T and Sτ , gives a first impression of the FIG. 1. Scatter plots of conformations of the (a)-(d) semiflexible (bending restrained) and (e)-(h) flexible (bend- ing unrestrained) polymers with 40 monomers in (q1, q2) space. Lightgray regions represent the generalized ensemble of all conformations found at all temperatures T and torsion strengths Sτ simulated. Black regions correspond to the most populated states at given Sτ values. Representative confor- mations are shown. Distances are measured in units of the length scale r0, given by the location of the LJ potential minimum. The van der Waals radius of a monomer is chosen to be σ = 2−1/6r0. The FENE (finitely extensible nonlin- ear elastic) bond potential is employed in the form [25] vbond(r) = log{1 − [(r − r0)/R]2}. The bond strength is sr = −KR2/2ǫ; the parameters were set to standard 3 system. The space of macrostates that share a charac- teristic structural feature such as the number of helical segments in a helix bundle forms a structural phase. Transitions temperatures for the various model param- eter settings were identified by standard canonical anal- yses of extremal fluctuations of energy (specific heat), structural quantities (e.g., radius of gyration), and order parameters (hq1i, hq2i, number of helices). This will be discussed in more detail elsewhere [33]. Based on this in- formation, the (q1, q2) space can be separated into regions ("structural phases") as shown in Fig. 3 for bending- restrained semiflexible (left figure) and unrestrained, flex- ible polymers (right figure). Black lines represent folding trajectories for several single polymers with fixed torsion strengths in the interval Sτ ∈ [0, 30] in (q1, q2) space upon cooling. All trajectories begin in the high-temperature, random-coil phase (upper right corner in Fig. 3, i.e., large q1, q2 values) and propagate toward a helical state by de- creasing the temperature. The folding channels at given Sτ values effectively connect free-energy minima (dots) at various temperatures. The structural phases of flexible polymers are less well separated, in which case folding channels, after passing a liquid phase, end in the solid amorphous phase. This general transition behavior is virtually independent of the torsion strength Sτ . However, the influence of the value of Sτ upon helix and helix-bundle formation is sig- nificant for the bending-restrained, semiflexible polymer. For Sτ = 0, the behavior is similar to that of the flexible polymer, but for torsion strength Sτ = 1, it crystallizes initially into a three-helix bundle and then undergoes a solid-solid transition. It emerges from it as a four-helix bundle, which is energetically slightly more favorable at very low temperatures. Three-helix bundles clearly form at sufficiently large torsion strength (e.g., for Sτ = 5). Increasing the torsion strength further favors the ex- tension of helical segments, compared to local nonbonded contacts. The torsional interaction overcompensates what had been an energetic gain of nonbonded monomer- monomer contacts despite necessary bending penalties. The number of turns is reduced to a single one and a double-helix forms in the solid phase (Sτ = 8). Bending- restrained polymers with Sτ = 20 coexist in a transition state between single- and double helix, i.e., in an inter- mediate ordered solid phase a single helix is formed first, which upon further cooling splits into a double-helix at the expense of the formation of a single turn. This torsion strength marks the threshold at which distant monomer- monomer contacts are still formed. For torsion strengths close to Sτ = 30 and beyond, only stable single-helix phases form. The complete structural hyper-phase diagram, parametrized by temperature T and torsion strength Sτ , is depicted in Fig. 4 for bending-restrained, semi- flexible polymers (left) and for unrestrained, flexible polymers with torsion (right). The phase diagram FIG. 2. Ratio of order parameters hq2i/hq1i for lowest-energy structures at various values of Sτ in the cases of restrained and unrestrained bending. differences of the conformational phases of entire classes of semiflexible and flexible polymers. It depicts the pos- sible folding channels for the polymers. From the fig- ure, it is obvious that the distributions spread out much more for the bending-restrained polymer. Individual sec- tions (phases) are clearly separated, with less-populated regions in-between. This is different in the case of flex- ible polymers. Although conformational phases can be identified as well, their separation is much less promi- nent. These differences can be interpreted in the way that in the case of semiflexible helical polymers, structural phases are more stable, because these are surrounded by entropically suppressed regions, which cause free-energy barriers and phase separation between the helical phases. The black regions in the figures represent the populations in (q1, q2) for fixed Sτ values, i.e., for individual polymer systems and confirm that semiflexible polymers with a certain torsion strength prefer to form tertiary structures inside a distinct helical phase only, which is not the case for flexible polymers. The differences in their structural behavior are also clearly visible when plotting the order parameter ratio hq2i/hq1i for the lowest-energy structures found in the simulations, as shown in Fig. 2. The step-like decrease of this ratio for the bending-restrained polymers enables the location of the threshold values of Sτ where structural phases are separated. 1, q′ For a more systematic analysis of the folding behav- ior and its dependence on the torsion strength Sτ and temperature T , we define the free-energy in order pa- rameter space by FSτ ,T (q1, q2) = −kBT log ZSτ ,T (q1, q2), where ZSτ ,T (q′ 2 − q2(X)) exp[−E(X)/kBT ] is the restricted partition func- tion in the space of all polymer conformations {X}. Fix- ing Sτ and T , the free energy FSτ ,T (q1, q2) possesses a global minimum at order parameter values (qmin ). The ensemble of all conformations with these order pa- rameter values represents a dominant macrostate of the 2) = R DXδ(q′ 1 − q1(X))δ(q′ , qmin 1 2 4 FIG. 3. Structural phase diagrams for bending-restrained semiflexible (left) and unrestrained flexible polymers (right) in (q1, q2) order parameter space for the temperature and torsion strength space (T, Sτ ) covered in our simulations. Colored regions represent structural phases. Black dots locate free energy minima at given T and Sτ values. Trajectories show the helical folding pathways at fixed torsion strengths Sτ by decreasing the temperature. for the semiflexible polymer exhibits apparently more structure in the folded regime at temperatures T < 0.5 over the entire interval of torsion strengths. Whereas at torsion strengths Sτ < 7 four-helix bundles, three-helix bundles, and amorphous conformations compete and the phases sensitively depend on the temperature, two-helix bundles and single-helix conformations are clearly dominant for Sτ > 7. Remarkably, the liquid (globular) phase disappears for sufficiently large torsion strengths (Sτ > 15), in which case direct coil-helix transitions 30 25 20  S 15 10 4h 5 0 1h 2h 3h L C C A L 0.2 0.4 0.6 A 0.8 T 1.0 1.2 1.4 0.2 0.4 0.6 1.0 1.2 1.4 0.8 T FIG. 4. Hyper-phase diagrams of bending-restrained semi- flexible (left) and unrestrained flexible polymers (right) with 40 monomers, represented in the space of the torsion strength Sτ as a material parameter distinguishing classes of polymers and the temperature T as an external control parameter for the formation of structural phases. The color code is the same as in Fig. 3. occur. Within the range 15 < Sτ < 27, helix-helix (solid-solid) transitions are present, where single helices collapse into two-helix bundles by forming a turn. Once the torsion strength dominates over nonbonded monomer-monomer interaction, i.e., for Sτ > 27, the well-known direct transition from random coils into single helices occurs. Contrarily, the folding process of flexible polymers is hardly affected qualitatively by torsional constraints [Fig. 4 (right)]. The three phases of random coils, glob- ular, and amorphous structures are well separate, but a helical phase is nonexistent. Furthermore, if bending is not restrained, the liquid phase does not disappear and thus a helix-coil transition does not occur. In this Letter, we have systematically investigated the influence of bending restraints upon the formation of sta- ble helical phases. We determined all structural phases for entire classes of flexible and semiflexible polymers with torsion. These results were summarized in struc- tural hyperphase diagrams for both polymer classes. The primary result our study is that an effective bend- ing restraint along the polymer chain is necessary to sta- bilize helical structures and, in particular, helix bundles. Different helical structure types that are separated by entropic gaps in conformational space can only be identi- fied clearly for semiflexible polymers, whereas for flexible polymers torsional barriers alone are not sufficient to sta- bilize individual helical phases. The outcome of this study provides evidence for the natural preference and significance of locally ordered he- lical secondary structures for semiflexible biopolymers, which effectively include DNA and most proteins. Our results support the understanding of the almost strict confinement of bond angles in polypeptides (such as bio- proteins), which reduces the set of degrees of freedom that participate in their functional structure formation to dihedral angles. For this reason, it is unlikely that flex- ible polymers, i.e., polymers without bending restraint, can be vital and functional in a biological system. This work has been supported partially by the NSF under Grant No. DMR-1207437 and by CNPq (National Council for Scientific and Technological Development, Brazil) under Grant No. 402091/2012-4. ∗ E-mail: [email protected] † E-mail: [email protected]; Homepage: http://www.smsyslab.org 5 (2005). [12] S. A. Sabeur, Cent. Eur. J. Phys. 12, 421 (2014). [13] S. R. Presnell and F. E. Cohen, Proc. Natl. Acad. Sci. 86, 6592 (1989). [14] Z. Guo, C. L. Brooks III, and E. M. Boczko, Proc. Natl. Acad. Sci. 94, 10161 (1997). [15] C. Zhang, J. Hou, and S. H. Kim, Proc. Natl. Acad. Sci. 99, 3581 (2002). [16] S. W. Bruun, V. Iesmantavicius, J. Danielsson, and F. M. Poulsen, Proc. Natl. Acad. Sci. 107, 13306 (2010). [17] T. Bereau, M. Bachmann, and M. Deserno, J. Am. Chem. Soc. 132, 13129 (2010). [18] T. Bereau, M. Deserno, and M. Bachmann, Biophys. J. 100, 2764 (2011). [19] P. Palenc´ar and T. Bleha, Comput. Theor. Chem. 1006, 62 (2013). [20] Z. Qin, A. Fabre, and M. J. Buehler, Eur. Phys. J. E 36, 53 (2013). [21] G. C. Rollins and K. A. Dill, J. Am. Chem. Soc. 136, 11420 (2014). [22] S. Schnabel, T. Vogel, M. Bachmann, and W. Janke, Chem. Phys. Lett. 476, 201 (2009). [23] S. Schnabel, M. Bachmann, and W. Janke, J. Chem. [1] A. Maritan, C. Micheletti, A. Trovato, and J. R. Banavar, Phys. 131, 124904 (2009). Nature 406, 287 (2000). [24] D. T. Seaton, S. Schnabel, D. P. Landau, and M. Bach- [2] T. Vogel, T. Neuhaus, M. Bachmann, and W. Janke, mann, Phys. Rev. Lett. 110, 028103 (2013). Europhys. Lett. 85, 10003 (2009). [25] K. Kremer and G. S. Grest, J. Chem. Phys. 92, 5057 [3] T. Vogel, T. Neuhaus, M. Bachmann, and W. Janke, (1990). Phys. Rev. E 80, 011802 (2009). [26] R. H. Swendsen and J. S. Wang, Phys. Rev. Lett. 57, [4] B. Zimm and J. Bragg, J. Chem. Phys. 28, 1246 (1958). [5] B. Zimm and J. Bragg, J. Chem. Phys. 31, 526 (1959). [6] D. Poland and H. A. Scheraga, Theory of Helix-Coil Transitions in Biopolymers (Academic Press, New York, 1970). [7] A. V. Badasyan, A. Giacometti, Y. Sh. Mamasakhlisov, V. F. Morozov, and A. S. Benight, Phys. Rev. E 81, 021921 (2010). [8] M. Bachmann, Thermodynamics and Statistical Mechan- ics of Macromolecular Systems (Cambridge University Press, Cambridge, 2014). 2607 (1986). [27] K. Hukushima and K. Nemoto, J. Phys. Soc. Jpn. 65, 1604 (1996). [28] K. Hukushima, H. Takayama, and K. Nemoto, Int. J. Mod. Phys. C 7, 337 (1996). [29] C. J. Geyer, Comp. Sci. Stat. 54, 156 (1991). [30] S. Schnabel, M. Bachmann, and W. Janke, Phys. Rev. Lett. 98, 048103 (2007). [31] C. Junghans, M. Bachmann, and W. Janke, Phys. Rev. Lett. 97, 218103 (2006). [32] C. Junghans, M. Bachmann, and W. Janke, J. Chem. [9] J. P. Kemp and Z. Y. Chen, Phys. Rev. Lett. 81, 3880 Phys. 128, 085103 (2008). (1998). [33] M. J. Williams and M. Bachmann, to be published [10] D. C. Rapaport, Phys. Rev. E 66, 011906 (2002). [11] V. Varshney and G. A. Carri, Phys. Rev. Lett. 95, 168304 (2015).
1809.03472
1
1809
2018-09-10T17:44:20
Spatial control of irreversible protein aggregation
[ "physics.bio-ph", "cond-mat.soft", "q-bio.SC" ]
Liquid cellular compartments spatially segregate from the cytoplasm and can regulate aberrant protein aggregation, a process linked to several medical conditions, including Alzheimer's and Parkinson's diseases. Yet the mechanisms by which these droplet-like compartments affect protein aggregation remain unknown. Here, we combine kinetic theory of protein aggregation and liquid-liquid phase separation to study the spatial control of irreversible protein aggregation in the presence of liquid compartments. We find that, even for weak interactions between the compartment constituents and the aggregating monomers, aggregates are strongly enriched inside the liquid compartment relative to the surrounding cytoplasm. We show that this enrichment is caused by a positive feedback mechanism of aggregate nucleation and growth which is mediated by a flux maintaining the phase equilibrium between the compartment and the cytoplasm. Our model predicts that the compartment volume that maximizes aggregate enrichment in the compartment is determined by the reaction orders of aggregate nucleation. The underlying mechanism of aggregate enrichment could be used to confine cytotoxic protein aggregates inside droplet-like compartments suggesting potential new avenues against aberrant protein aggregation. Our findings could also represent a common mechanism for the spatial control of irreversible chemical reactions in general.
physics.bio-ph
physics
Spatial control of irreversible protein aggregation Christoph A. Weber,1, 2 Thomas C. T. Michaels,1, 2 and L. Mahadevan3 1Engineering and Applied Sciences, Cambridge, Massachusetts 02138, United States of America 3Engineering and Applied Sciences, Physics and Organismic and Evolutionary Biology, Cambridge, Massachusetts 02138, United States of America 2contributed equally Liquid cellular compartments spatially segregate from the cytoplasm and can regulate aberrant protein aggregation, a process linked to several medical conditions, including Alzheimer's and Parkinson's diseases. Yet the mechanisms by which these droplet-like compartments affect pro- tein aggregation remain unknown. Here, we combine kinetic theory of protein aggregation and liquid-liquid phase separation to study the spatial control of irreversible protein aggregation in the presence of liquid compartments. We find that, even for weak interactions between the compart- ment constituents and the aggregating monomers, aggregates are strongly enriched inside the liquid compartment relative to the surrounding cytoplasm. We show that this enrichment is caused by a positive feedback mechanism of aggregate nucleation and growth which is mediated by a flux maintaining the phase equilibrium between the compartment and the cytoplasm. Our model pre- dicts that the compartment volume that maximizes aggregate enrichment in the compartment is determined by the reaction orders of aggregate nucleation. The underlying mechanism of aggregate enrichment could be used to confine cytotoxic protein aggregates inside droplet-like compartments suggesting potential new avenues against aberrant protein aggregation. Our findings could also represent a common mechanism for the spatial control of irreversible chemical reactions in general. Spatial control within living cells is essential to many cellular activities, ranging from the local control of pro- tein activity to the uptake of pathogens or the manage- ment of wastes [1]. Understanding the mechanisms un- derlying regulation of cell activities in space and time is key not only for biological function, but also in view of understanding and eventually controlling cellular dys- function [2 -- 5]. The spatial organization of cellular activ- ities is often associated with membrane-bound organelles that ensure permeation only for certain molecules of specific molecular structure [6 -- 8]. Recently, new types of organelles have been discovered that do not possess a membrane. They are referred to as non-membrane bound compartments and they share most hallmark properties with actual liquid-like droplets [9 -- 13]. Un- like organelles surrounded by membranes, these non- membrane bound compartments are formed by liquid- liquid phase separation. In many cases this phase sepa- ration is driven by disfavouring interactions between the constituent molecules of the compartment and the sur- rounding cytoplasm [14, 15]. The partitioning of other in- tracellular molecules into such droplet-like compartments is then controlled by their relative interactions with the constituent molecules of the compartment. These droplet-like compartments are ubiquitous in- side living cells [13]. For instance, they emerge prior to cell division [9, 16], and form as a response to cellu- lar stress [17 -- 19]. They have been shown to enrich pro- teins [20 -- 22] and genetic material [16, 23, 24] providing distinct environments for chemical reactions and biologi- cal function. The molecules hosted inside these compart- ments may even be protected against other agents from the cytoplasm [25] or face conditions facilitating their molecular repair [22, 26 -- 30]. In addition to these roles, recent evidence suggests that liquid cellular compart- FIG. 1. Enrichment of monomers and aggregates via liquid-like compartments. Protein aggregation may occur homogeneously inside cells also leading to aggregates inside more sensitive cellular regions (left). A liquid com- partment may accumulate monomers and thereby trigger the local formation of aggregates (right). The hardly diffusing aggregates are thus kept away from a more sensible cellular region. Such a spatial segregation of aggregates is ideal for adding functional, drug-like molecules which dominantly dis- solve inside the compartment. These molecules may degrade the aggregates or inhibit further growth and nucleation. But most importantly, as these molecules are localized inside the compartment their toxic effects are diminished. ments could play an important role in regulating patho- logical protein aggregation [31, 32]. An example is the ir- reversible assembly of amyloids into fibrillar aggregates, a process that is linked to a large variety of currently incur- able diseases [2, 33 -- 37], such as Alzheimer's and Parkin- son's diseases, amyloidosis or type-II diabetes. As an- other example, a chaperone in yeast uses a prion-like, in- trinsically disordered domain to bind and sequester mis- folded proteins in protein deposition sites [38, 39]. More- over, misfolded and pathological proteins can accumulate 8 1 0 2 p e S 0 1 ] h p - o i b . s c i s y h p [ 1 v 2 7 4 3 0 . 9 0 8 1 : v i X r a IIImonomeraggregateliquidcompartment<latexit sha1_base64="vMCmCBvrQqvMkcZe5c78yK19qu0=">AAAB63icbVBNSwMxEJ3Ur1q/qh69BIvgqeyKoMeiF48VbC20S8mm2W5okl2SrFCW/gUvHhTx6h/y5r8x2+5BWx8MPN6bYWZemApurOd9o8ra+sbmVnW7trO7t39QPzzqmiTTlHVoIhLdC4lhgivWsdwK1ks1IzIU7DGc3Bb+4xPThifqwU5TFkgyVjzilNhCGqQxH9YbXtObA68SvyQNKNEe1r8Go4RmkilLBTGm73upDXKiLaeCzWqDzLCU0AkZs76jikhmgnx+6wyfOWWEo0S7UhbP1d8TOZHGTGXoOiWxsVn2CvE/r5/Z6DrIuUozyxRdLIoygW2Ci8fxiGtGrZg6Qqjm7lZMY6IJtS6emgvBX355lXQvmr7X9O8vG62bMo4qnMApnIMPV9CCO2hDByjE8Ayv8IYkekHv6GPRWkHlzDH8Afr8AROIjj8=</latexit><latexit sha1_base64="vMCmCBvrQqvMkcZe5c78yK19qu0=">AAAB63icbVBNSwMxEJ3Ur1q/qh69BIvgqeyKoMeiF48VbC20S8mm2W5okl2SrFCW/gUvHhTx6h/y5r8x2+5BWx8MPN6bYWZemApurOd9o8ra+sbmVnW7trO7t39QPzzqmiTTlHVoIhLdC4lhgivWsdwK1ks1IzIU7DGc3Bb+4xPThifqwU5TFkgyVjzilNhCGqQxH9YbXtObA68SvyQNKNEe1r8Go4RmkilLBTGm73upDXKiLaeCzWqDzLCU0AkZs76jikhmgnx+6wyfOWWEo0S7UhbP1d8TOZHGTGXoOiWxsVn2CvE/r5/Z6DrIuUozyxRdLIoygW2Ci8fxiGtGrZg6Qqjm7lZMY6IJtS6emgvBX355lXQvmr7X9O8vG62bMo4qnMApnIMPV9CCO2hDByjE8Ayv8IYkekHv6GPRWkHlzDH8Afr8AROIjj8=</latexit><latexit sha1_base64="vMCmCBvrQqvMkcZe5c78yK19qu0=">AAAB63icbVBNSwMxEJ3Ur1q/qh69BIvgqeyKoMeiF48VbC20S8mm2W5okl2SrFCW/gUvHhTx6h/y5r8x2+5BWx8MPN6bYWZemApurOd9o8ra+sbmVnW7trO7t39QPzzqmiTTlHVoIhLdC4lhgivWsdwK1ks1IzIU7DGc3Bb+4xPThifqwU5TFkgyVjzilNhCGqQxH9YbXtObA68SvyQNKNEe1r8Go4RmkilLBTGm73upDXKiLaeCzWqDzLCU0AkZs76jikhmgnx+6wyfOWWEo0S7UhbP1d8TOZHGTGXoOiWxsVn2CvE/r5/Z6DrIuUozyxRdLIoygW2Ci8fxiGtGrZg6Qqjm7lZMY6IJtS6emgvBX355lXQvmr7X9O8vG62bMo4qnMApnIMPV9CCO2hDByjE8Ayv8IYkekHv6GPRWkHlzDH8Afr8AROIjj8=</latexit><latexit sha1_base64="vMCmCBvrQqvMkcZe5c78yK19qu0=">AAAB63icbVBNSwMxEJ3Ur1q/qh69BIvgqeyKoMeiF48VbC20S8mm2W5okl2SrFCW/gUvHhTx6h/y5r8x2+5BWx8MPN6bYWZemApurOd9o8ra+sbmVnW7trO7t39QPzzqmiTTlHVoIhLdC4lhgivWsdwK1ks1IzIU7DGc3Bb+4xPThifqwU5TFkgyVjzilNhCGqQxH9YbXtObA68SvyQNKNEe1r8Go4RmkilLBTGm73upDXKiLaeCzWqDzLCU0AkZs76jikhmgnx+6wyfOWWEo0S7UhbP1d8TOZHGTGXoOiWxsVn2CvE/r5/Z6DrIuUozyxRdLIoygW2Ci8fxiGtGrZg6Qqjm7lZMY6IJtS6emgvBX355lXQvmr7X9O8vG62bMo4qnMApnIMPV9CCO2hDByjE8Ayv8IYkekHv6GPRWkHlzDH8Afr8AROIjj8=</latexit>,a<latexit sha1_base64="K3YBRu8n4+INWsi7VwmBsJ8SQaw=">AAAB83icbVBNS8NAEJ3Ur1q/qh69LBbBg5REBD0WvXisYGuhCWWz3bRLd5OwOxFL6N/w4kERr/4Zb/4bt20O2vpg4PHeDDPzwlQKg6777ZRWVtfWN8qbla3tnd296v5B2ySZZrzFEpnoTkgNlyLmLRQoeSfVnKpQ8odwdDP1Hx65NiKJ73Gc8kDRQSwiwShayT8jPvIn1Cqnk1615tbdGcgy8QpSgwLNXvXL7ycsUzxGJqkxXc9NMcipRsEkn1T8zPCUshEd8K6lMVXcBPns5gk5sUqfRIm2FSOZqb8ncqqMGavQdiqKQ7PoTcX/vG6G0VWQizjNkMdsvijKJMGETAMgfaE5Qzm2hDIt7K2EDammDG1MFRuCt/jyMmmf1z237t1d1BrXRRxlOIJjOAUPLqEBt9CEFjBI4Rle4c3JnBfn3fmYt5acYuYQ/sD5/AHnXJGX</latexit><latexit sha1_base64="K3YBRu8n4+INWsi7VwmBsJ8SQaw=">AAAB83icbVBNS8NAEJ3Ur1q/qh69LBbBg5REBD0WvXisYGuhCWWz3bRLd5OwOxFL6N/w4kERr/4Zb/4bt20O2vpg4PHeDDPzwlQKg6777ZRWVtfWN8qbla3tnd296v5B2ySZZrzFEpnoTkgNlyLmLRQoeSfVnKpQ8odwdDP1Hx65NiKJ73Gc8kDRQSwiwShayT8jPvIn1Cqnk1615tbdGcgy8QpSgwLNXvXL7ycsUzxGJqkxXc9NMcipRsEkn1T8zPCUshEd8K6lMVXcBPns5gk5sUqfRIm2FSOZqb8ncqqMGavQdiqKQ7PoTcX/vG6G0VWQizjNkMdsvijKJMGETAMgfaE5Qzm2hDIt7K2EDammDG1MFRuCt/jyMmmf1z237t1d1BrXRRxlOIJjOAUPLqEBt9CEFjBI4Rle4c3JnBfn3fmYt5acYuYQ/sD5/AHnXJGX</latexit><latexit sha1_base64="K3YBRu8n4+INWsi7VwmBsJ8SQaw=">AAAB83icbVBNS8NAEJ3Ur1q/qh69LBbBg5REBD0WvXisYGuhCWWz3bRLd5OwOxFL6N/w4kERr/4Zb/4bt20O2vpg4PHeDDPzwlQKg6777ZRWVtfWN8qbla3tnd296v5B2ySZZrzFEpnoTkgNlyLmLRQoeSfVnKpQ8odwdDP1Hx65NiKJ73Gc8kDRQSwiwShayT8jPvIn1Cqnk1615tbdGcgy8QpSgwLNXvXL7ycsUzxGJqkxXc9NMcipRsEkn1T8zPCUshEd8K6lMVXcBPns5gk5sUqfRIm2FSOZqb8ncqqMGavQdiqKQ7PoTcX/vG6G0VWQizjNkMdsvijKJMGETAMgfaE5Qzm2hDIt7K2EDammDG1MFRuCt/jyMmmf1z237t1d1BrXRRxlOIJjOAUPLqEBt9CEFjBI4Rle4c3JnBfn3fmYt5acYuYQ/sD5/AHnXJGX</latexit><latexit sha1_base64="K3YBRu8n4+INWsi7VwmBsJ8SQaw=">AAAB83icbVBNS8NAEJ3Ur1q/qh69LBbBg5REBD0WvXisYGuhCWWz3bRLd5OwOxFL6N/w4kERr/4Zb/4bt20O2vpg4PHeDDPzwlQKg6777ZRWVtfWN8qbla3tnd296v5B2ySZZrzFEpnoTkgNlyLmLRQoeSfVnKpQ8odwdDP1Hx65NiKJ73Gc8kDRQSwiwShayT8jPvIn1Cqnk1615tbdGcgy8QpSgwLNXvXL7ycsUzxGJqkxXc9NMcipRsEkn1T8zPCUshEd8K6lMVXcBPns5gk5sUqfRIm2FSOZqb8ncqqMGavQdiqKQ7PoTcX/vG6G0VWQizjNkMdsvijKJMGETAMgfaE5Qzm2hDIt7K2EDammDG1MFRuCt/jyMmmf1z237t1d1BrXRRxlOIJjOAUPLqEBt9CEFjBI4Rle4c3JnBfn3fmYt5acYuYQ/sD5/AHnXJGX</latexit>,m<latexit sha1_base64="NRD1qlY6xOZWNfsOVknrXjMAT00=">AAAB83icbVBNS8NAEJ3Ur1q/qh69LBbBg5REBD0WvXisYGuhCWWz3bRLd5OwOxFL6N/w4kERr/4Zb/4bt20O2vpg4PHeDDPzwlQKg6777ZRWVtfWN8qbla3tnd296v5B2ySZZrzFEpnoTkgNlyLmLRQoeSfVnKpQ8odwdDP1Hx65NiKJ73Gc8kDRQSwiwShayT8jPvIn1CpXk1615tbdGcgy8QpSgwLNXvXL7ycsUzxGJqkxXc9NMcipRsEkn1T8zPCUshEd8K6lMVXcBPns5gk5sUqfRIm2FSOZqb8ncqqMGavQdiqKQ7PoTcX/vG6G0VWQizjNkMdsvijKJMGETAMgfaE5Qzm2hDIt7K2EDammDG1MFRuCt/jyMmmf1z237t1d1BrXRRxlOIJjOAUPLqEBt9CEFjBI4Rle4c3JnBfn3fmYt5acYuYQ/sD5/AH5mJGj</latexit><latexit sha1_base64="NRD1qlY6xOZWNfsOVknrXjMAT00=">AAAB83icbVBNS8NAEJ3Ur1q/qh69LBbBg5REBD0WvXisYGuhCWWz3bRLd5OwOxFL6N/w4kERr/4Zb/4bt20O2vpg4PHeDDPzwlQKg6777ZRWVtfWN8qbla3tnd296v5B2ySZZrzFEpnoTkgNlyLmLRQoeSfVnKpQ8odwdDP1Hx65NiKJ73Gc8kDRQSwiwShayT8jPvIn1CpXk1615tbdGcgy8QpSgwLNXvXL7ycsUzxGJqkxXc9NMcipRsEkn1T8zPCUshEd8K6lMVXcBPns5gk5sUqfRIm2FSOZqb8ncqqMGavQdiqKQ7PoTcX/vG6G0VWQizjNkMdsvijKJMGETAMgfaE5Qzm2hDIt7K2EDammDG1MFRuCt/jyMmmf1z237t1d1BrXRRxlOIJjOAUPLqEBt9CEFjBI4Rle4c3JnBfn3fmYt5acYuYQ/sD5/AH5mJGj</latexit><latexit sha1_base64="NRD1qlY6xOZWNfsOVknrXjMAT00=">AAAB83icbVBNS8NAEJ3Ur1q/qh69LBbBg5REBD0WvXisYGuhCWWz3bRLd5OwOxFL6N/w4kERr/4Zb/4bt20O2vpg4PHeDDPzwlQKg6777ZRWVtfWN8qbla3tnd296v5B2ySZZrzFEpnoTkgNlyLmLRQoeSfVnKpQ8odwdDP1Hx65NiKJ73Gc8kDRQSwiwShayT8jPvIn1CpXk1615tbdGcgy8QpSgwLNXvXL7ycsUzxGJqkxXc9NMcipRsEkn1T8zPCUshEd8K6lMVXcBPns5gk5sUqfRIm2FSOZqb8ncqqMGavQdiqKQ7PoTcX/vG6G0VWQizjNkMdsvijKJMGETAMgfaE5Qzm2hDIt7K2EDammDG1MFRuCt/jyMmmf1z237t1d1BrXRRxlOIJjOAUPLqEBt9CEFjBI4Rle4c3JnBfn3fmYt5acYuYQ/sD5/AH5mJGj</latexit><latexit sha1_base64="NRD1qlY6xOZWNfsOVknrXjMAT00=">AAAB83icbVBNS8NAEJ3Ur1q/qh69LBbBg5REBD0WvXisYGuhCWWz3bRLd5OwOxFL6N/w4kERr/4Zb/4bt20O2vpg4PHeDDPzwlQKg6777ZRWVtfWN8qbla3tnd296v5B2ySZZrzFEpnoTkgNlyLmLRQoeSfVnKpQ8odwdDP1Hx65NiKJ73Gc8kDRQSwiwShayT8jPvIn1CpXk1615tbdGcgy8QpSgwLNXvXL7ycsUzxGJqkxXc9NMcipRsEkn1T8zPCUshEd8K6lMVXcBPns5gk5sUqfRIm2FSOZqb8ncqqMGavQdiqKQ7PoTcX/vG6G0VWQizjNkMdsvijKJMGETAMgfaE5Qzm2hDIt7K2EDammDG1MFRuCt/jyMmmf1z237t1d1BrXRRxlOIJjOAUPLqEBt9CEFjBI4Rle4c3JnBfn3fmYt5acYuYQ/sD5/AH5mJGj</latexit>I<latexit sha1_base64="uj1Lu3/A7RVg+FAHZIJDy/GRYGo=">AAAB+HicbVBNSwMxEM3Wr1o/uurRS7AInsquCHosetFbBfsB7VqyabYNTbJLMivWpb/EiwdFvPpTvPlvTNs9aOuDgcd7M8zMCxPBDXjet1NYWV1b3yhulra2d3bL7t5+08SppqxBYxHrdkgME1yxBnAQrJ1oRmQoWCscXU391gPThsfqDsYJCyQZKB5xSsBKPbfcTYb8vgvsEbTMbiY9t+JVvRnwMvFzUkE56j33q9uPaSqZAiqIMR3fSyDIiAZOBZuUuqlhCaEjMmAdSxWRzATZ7PAJPrZKH0extqUAz9TfExmRxoxlaDslgaFZ9Kbif14nhegiyLhKUmCKzhdFqcAQ42kKuM81oyDGlhCqub0V0yHRhILNqmRD8BdfXibN06rvVf3bs0rtMo+jiA7RETpBPjpHNXSN6qiBKErRM3pFb86T8+K8Ox/z1oKTzxygP3A+fwBLqJN9</latexit><latexit sha1_base64="uj1Lu3/A7RVg+FAHZIJDy/GRYGo=">AAAB+HicbVBNSwMxEM3Wr1o/uurRS7AInsquCHosetFbBfsB7VqyabYNTbJLMivWpb/EiwdFvPpTvPlvTNs9aOuDgcd7M8zMCxPBDXjet1NYWV1b3yhulra2d3bL7t5+08SppqxBYxHrdkgME1yxBnAQrJ1oRmQoWCscXU391gPThsfqDsYJCyQZKB5xSsBKPbfcTYb8vgvsEbTMbiY9t+JVvRnwMvFzUkE56j33q9uPaSqZAiqIMR3fSyDIiAZOBZuUuqlhCaEjMmAdSxWRzATZ7PAJPrZKH0extqUAz9TfExmRxoxlaDslgaFZ9Kbif14nhegiyLhKUmCKzhdFqcAQ42kKuM81oyDGlhCqub0V0yHRhILNqmRD8BdfXibN06rvVf3bs0rtMo+jiA7RETpBPjpHNXSN6qiBKErRM3pFb86T8+K8Ox/z1oKTzxygP3A+fwBLqJN9</latexit><latexit sha1_base64="uj1Lu3/A7RVg+FAHZIJDy/GRYGo=">AAAB+HicbVBNSwMxEM3Wr1o/uurRS7AInsquCHosetFbBfsB7VqyabYNTbJLMivWpb/EiwdFvPpTvPlvTNs9aOuDgcd7M8zMCxPBDXjet1NYWV1b3yhulra2d3bL7t5+08SppqxBYxHrdkgME1yxBnAQrJ1oRmQoWCscXU391gPThsfqDsYJCyQZKB5xSsBKPbfcTYb8vgvsEbTMbiY9t+JVvRnwMvFzUkE56j33q9uPaSqZAiqIMR3fSyDIiAZOBZuUuqlhCaEjMmAdSxWRzATZ7PAJPrZKH0extqUAz9TfExmRxoxlaDslgaFZ9Kbif14nhegiyLhKUmCKzhdFqcAQ42kKuM81oyDGlhCqub0V0yHRhILNqmRD8BdfXibN06rvVf3bs0rtMo+jiA7RETpBPjpHNXSN6qiBKErRM3pFb86T8+K8Ox/z1oKTzxygP3A+fwBLqJN9</latexit><latexit sha1_base64="uj1Lu3/A7RVg+FAHZIJDy/GRYGo=">AAAB+HicbVBNSwMxEM3Wr1o/uurRS7AInsquCHosetFbBfsB7VqyabYNTbJLMivWpb/EiwdFvPpTvPlvTNs9aOuDgcd7M8zMCxPBDXjet1NYWV1b3yhulra2d3bL7t5+08SppqxBYxHrdkgME1yxBnAQrJ1oRmQoWCscXU391gPThsfqDsYJCyQZKB5xSsBKPbfcTYb8vgvsEbTMbiY9t+JVvRnwMvFzUkE56j33q9uPaSqZAiqIMR3fSyDIiAZOBZuUuqlhCaEjMmAdSxWRzATZ7PAJPrZKH0extqUAz9TfExmRxoxlaDslgaFZ9Kbif14nhegiyLhKUmCKzhdFqcAQ42kKuM81oyDGlhCqub0V0yHRhILNqmRD8BdfXibN06rvVf3bs0rtMo+jiA7RETpBPjpHNXSN6qiBKErRM3pFb86T8+K8Ox/z1oKTzxygP3A+fwBLqJN9</latexit>II<latexit sha1_base64="Q+wlYgKtwjfDzs+1/njkyYyv+wU=">AAAB+XicbVBNSwMxEM3Wr1q/Vj16CRbBU9kVQY9FL/ZWwX5Au5Zsmm1Dk+ySzBbL0n/ixYMiXv0n3vw3pu0etPXBwOO9GWbmhYngBjzv2ymsrW9sbhW3Szu7e/sH7uFR08SppqxBYxHrdkgME1yxBnAQrJ1oRmQoWCsc3c781phpw2P1AJOEBZIMFI84JWClnut2kyF/7AJ7Ai2zWm3ac8texZsDrxI/J2WUo95zv7r9mKaSKaCCGNPxvQSCjGjgVLBpqZsalhA6IgPWsVQRyUyQzS+f4jOr9HEUa1sK8Fz9PZERacxEhrZTEhiaZW8m/ud1Uoiug4yrJAWm6GJRlAoMMZ7FgPtcMwpiYgmhmttbMR0STSjYsEo2BH/55VXSvKj4XsW/vyxXb/I4iugEnaJz5KMrVEV3qI4aiKIxekav6M3JnBfn3flYtBacfOYY/YHz+QPlc5PQ</latexit><latexit sha1_base64="Q+wlYgKtwjfDzs+1/njkyYyv+wU=">AAAB+XicbVBNSwMxEM3Wr1q/Vj16CRbBU9kVQY9FL/ZWwX5Au5Zsmm1Dk+ySzBbL0n/ixYMiXv0n3vw3pu0etPXBwOO9GWbmhYngBjzv2ymsrW9sbhW3Szu7e/sH7uFR08SppqxBYxHrdkgME1yxBnAQrJ1oRmQoWCsc3c781phpw2P1AJOEBZIMFI84JWClnut2kyF/7AJ7Ai2zWm3ac8texZsDrxI/J2WUo95zv7r9mKaSKaCCGNPxvQSCjGjgVLBpqZsalhA6IgPWsVQRyUyQzS+f4jOr9HEUa1sK8Fz9PZERacxEhrZTEhiaZW8m/ud1Uoiug4yrJAWm6GJRlAoMMZ7FgPtcMwpiYgmhmttbMR0STSjYsEo2BH/55VXSvKj4XsW/vyxXb/I4iugEnaJz5KMrVEV3qI4aiKIxekav6M3JnBfn3flYtBacfOYY/YHz+QPlc5PQ</latexit><latexit sha1_base64="Q+wlYgKtwjfDzs+1/njkyYyv+wU=">AAAB+XicbVBNSwMxEM3Wr1q/Vj16CRbBU9kVQY9FL/ZWwX5Au5Zsmm1Dk+ySzBbL0n/ixYMiXv0n3vw3pu0etPXBwOO9GWbmhYngBjzv2ymsrW9sbhW3Szu7e/sH7uFR08SppqxBYxHrdkgME1yxBnAQrJ1oRmQoWCsc3c781phpw2P1AJOEBZIMFI84JWClnut2kyF/7AJ7Ai2zWm3ac8texZsDrxI/J2WUo95zv7r9mKaSKaCCGNPxvQSCjGjgVLBpqZsalhA6IgPWsVQRyUyQzS+f4jOr9HEUa1sK8Fz9PZERacxEhrZTEhiaZW8m/ud1Uoiug4yrJAWm6GJRlAoMMZ7FgPtcMwpiYgmhmttbMR0STSjYsEo2BH/55VXSvKj4XsW/vyxXb/I4iugEnaJz5KMrVEV3qI4aiKIxekav6M3JnBfn3flYtBacfOYY/YHz+QPlc5PQ</latexit><latexit sha1_base64="Q+wlYgKtwjfDzs+1/njkyYyv+wU=">AAAB+XicbVBNSwMxEM3Wr1q/Vj16CRbBU9kVQY9FL/ZWwX5Au5Zsmm1Dk+ySzBbL0n/ixYMiXv0n3vw3pu0etPXBwOO9GWbmhYngBjzv2ymsrW9sbhW3Szu7e/sH7uFR08SppqxBYxHrdkgME1yxBnAQrJ1oRmQoWCsc3c781phpw2P1AJOEBZIMFI84JWClnut2kyF/7AJ7Ai2zWm3ac8texZsDrxI/J2WUo95zv7r9mKaSKaCCGNPxvQSCjGjgVLBpqZsalhA6IgPWsVQRyUyQzS+f4jOr9HEUa1sK8Fz9PZERacxEhrZTEhiaZW8m/ud1Uoiug4yrJAWm6GJRlAoMMZ7FgPtcMwpiYgmhmttbMR0STSjYsEo2BH/55VXSvKj4XsW/vyxXb/I4iugEnaJz5KMrVEV3qI4aiKIxekav6M3JnBfn3flYtBacfOYY/YHz+QPlc5PQ</latexit>0<latexit sha1_base64="6ZTwbptvK00HUiMuNssEoeJJPkc=">AAAB6HicbVBNS8NAEJ3Ur1q/qh69LBbBU0lEqMeiF48t2A9oQ9lsJ+3azSbsboQS+gu8eFDEqz/Jm//GbZuDtj4YeLw3w8y8IBFcG9f9dgobm1vbO8Xd0t7+weFR+fikreNUMWyxWMSqG1CNgktsGW4EdhOFNAoEdoLJ3dzvPKHSPJYPZpqgH9GR5CFn1Fip6Q7KFbfqLkDWiZeTCuRoDMpf/WHM0gilYYJq3fPcxPgZVYYzgbNSP9WYUDahI+xZKmmE2s8Wh87IhVWGJIyVLWnIQv09kdFI62kU2M6ImrFe9ebif14vNeGNn3GZpAYlWy4KU0FMTOZfkyFXyIyYWkKZ4vZWwsZUUWZsNiUbgrf68jppX1U9t+o1ryv12zyOIpzBOVyCBzWowz00oAUMEJ7hFd6cR+fFeXc+lq0FJ585hT9wPn8AeRmMtA==</latexit><latexit sha1_base64="6ZTwbptvK00HUiMuNssEoeJJPkc=">AAAB6HicbVBNS8NAEJ3Ur1q/qh69LBbBU0lEqMeiF48t2A9oQ9lsJ+3azSbsboQS+gu8eFDEqz/Jm//GbZuDtj4YeLw3w8y8IBFcG9f9dgobm1vbO8Xd0t7+weFR+fikreNUMWyxWMSqG1CNgktsGW4EdhOFNAoEdoLJ3dzvPKHSPJYPZpqgH9GR5CFn1Fip6Q7KFbfqLkDWiZeTCuRoDMpf/WHM0gilYYJq3fPcxPgZVYYzgbNSP9WYUDahI+xZKmmE2s8Wh87IhVWGJIyVLWnIQv09kdFI62kU2M6ImrFe9ebif14vNeGNn3GZpAYlWy4KU0FMTOZfkyFXyIyYWkKZ4vZWwsZUUWZsNiUbgrf68jppX1U9t+o1ryv12zyOIpzBOVyCBzWowz00oAUMEJ7hFd6cR+fFeXc+lq0FJ585hT9wPn8AeRmMtA==</latexit><latexit sha1_base64="6ZTwbptvK00HUiMuNssEoeJJPkc=">AAAB6HicbVBNS8NAEJ3Ur1q/qh69LBbBU0lEqMeiF48t2A9oQ9lsJ+3azSbsboQS+gu8eFDEqz/Jm//GbZuDtj4YeLw3w8y8IBFcG9f9dgobm1vbO8Xd0t7+weFR+fikreNUMWyxWMSqG1CNgktsGW4EdhOFNAoEdoLJ3dzvPKHSPJYPZpqgH9GR5CFn1Fip6Q7KFbfqLkDWiZeTCuRoDMpf/WHM0gilYYJq3fPcxPgZVYYzgbNSP9WYUDahI+xZKmmE2s8Wh87IhVWGJIyVLWnIQv09kdFI62kU2M6ImrFe9ebif14vNeGNn3GZpAYlWy4KU0FMTOZfkyFXyIyYWkKZ4vZWwsZUUWZsNiUbgrf68jppX1U9t+o1ryv12zyOIpzBOVyCBzWowz00oAUMEJ7hFd6cR+fFeXc+lq0FJ585hT9wPn8AeRmMtA==</latexit><latexit sha1_base64="6ZTwbptvK00HUiMuNssEoeJJPkc=">AAAB6HicbVBNS8NAEJ3Ur1q/qh69LBbBU0lEqMeiF48t2A9oQ9lsJ+3azSbsboQS+gu8eFDEqz/Jm//GbZuDtj4YeLw3w8y8IBFcG9f9dgobm1vbO8Xd0t7+weFR+fikreNUMWyxWMSqG1CNgktsGW4EdhOFNAoEdoLJ3dzvPKHSPJYPZpqgH9GR5CFn1Fip6Q7KFbfqLkDWiZeTCuRoDMpf/WHM0gilYYJq3fPcxPgZVYYzgbNSP9WYUDahI+xZKmmE2s8Wh87IhVWGJIyVLWnIQv09kdFI62kU2M6ImrFe9ebif14vNeGNn3GZpAYlWy4KU0FMTOZfkyFXyIyYWkKZ4vZWwsZUUWZsNiUbgrf68jppX1U9t+o1ryv12zyOIpzBOVyCBzWowz00oAUMEJ7hFd6cR+fFeXc+lq0FJ585hT9wPn8AeRmMtA==</latexit>1<latexit sha1_base64="l9eImvYcFOKpzEDji/n9jPDeWb8=">AAAB6HicbVBNS8NAEJ3Ur1q/qh69LBbBU0lEqMeiF48t2A9oQ9lsJ+3azSbsboQS+gu8eFDEqz/Jm//GbZuDtj4YeLw3w8y8IBFcG9f9dgobm1vbO8Xd0t7+weFR+fikreNUMWyxWMSqG1CNgktsGW4EdhOFNAoEdoLJ3dzvPKHSPJYPZpqgH9GR5CFn1Fip6Q3KFbfqLkDWiZeTCuRoDMpf/WHM0gilYYJq3fPcxPgZVYYzgbNSP9WYUDahI+xZKmmE2s8Wh87IhVWGJIyVLWnIQv09kdFI62kU2M6ImrFe9ebif14vNeGNn3GZpAYlWy4KU0FMTOZfkyFXyIyYWkKZ4vZWwsZUUWZsNiUbgrf68jppX1U9t+o1ryv12zyOIpzBOVyCBzWowz00oAUMEJ7hFd6cR+fFeXc+lq0FJ585hT9wPn8Aep2MtQ==</latexit><latexit sha1_base64="l9eImvYcFOKpzEDji/n9jPDeWb8=">AAAB6HicbVBNS8NAEJ3Ur1q/qh69LBbBU0lEqMeiF48t2A9oQ9lsJ+3azSbsboQS+gu8eFDEqz/Jm//GbZuDtj4YeLw3w8y8IBFcG9f9dgobm1vbO8Xd0t7+weFR+fikreNUMWyxWMSqG1CNgktsGW4EdhOFNAoEdoLJ3dzvPKHSPJYPZpqgH9GR5CFn1Fip6Q3KFbfqLkDWiZeTCuRoDMpf/WHM0gilYYJq3fPcxPgZVYYzgbNSP9WYUDahI+xZKmmE2s8Wh87IhVWGJIyVLWnIQv09kdFI62kU2M6ImrFe9ebif14vNeGNn3GZpAYlWy4KU0FMTOZfkyFXyIyYWkKZ4vZWwsZUUWZsNiUbgrf68jppX1U9t+o1ryv12zyOIpzBOVyCBzWowz00oAUMEJ7hFd6cR+fFeXc+lq0FJ585hT9wPn8Aep2MtQ==</latexit><latexit sha1_base64="l9eImvYcFOKpzEDji/n9jPDeWb8=">AAAB6HicbVBNS8NAEJ3Ur1q/qh69LBbBU0lEqMeiF48t2A9oQ9lsJ+3azSbsboQS+gu8eFDEqz/Jm//GbZuDtj4YeLw3w8y8IBFcG9f9dgobm1vbO8Xd0t7+weFR+fikreNUMWyxWMSqG1CNgktsGW4EdhOFNAoEdoLJ3dzvPKHSPJYPZpqgH9GR5CFn1Fip6Q3KFbfqLkDWiZeTCuRoDMpf/WHM0gilYYJq3fPcxPgZVYYzgbNSP9WYUDahI+xZKmmE2s8Wh87IhVWGJIyVLWnIQv09kdFI62kU2M6ImrFe9ebif14vNeGNn3GZpAYlWy4KU0FMTOZfkyFXyIyYWkKZ4vZWwsZUUWZsNiUbgrf68jppX1U9t+o1ryv12zyOIpzBOVyCBzWowz00oAUMEJ7hFd6cR+fFeXc+lq0FJ585hT9wPn8Aep2MtQ==</latexit><latexit sha1_base64="l9eImvYcFOKpzEDji/n9jPDeWb8=">AAAB6HicbVBNS8NAEJ3Ur1q/qh69LBbBU0lEqMeiF48t2A9oQ9lsJ+3azSbsboQS+gu8eFDEqz/Jm//GbZuDtj4YeLw3w8y8IBFcG9f9dgobm1vbO8Xd0t7+weFR+fikreNUMWyxWMSqG1CNgktsGW4EdhOFNAoEdoLJ3dzvPKHSPJYPZpqgH9GR5CFn1Fip6Q3KFbfqLkDWiZeTCuRoDMpf/WHM0gilYYJq3fPcxPgZVYYzgbNSP9WYUDahI+xZKmmE2s8Wh87IhVWGJIyVLWnIQv09kdFI62kU2M6ImrFe9ebif14vNeGNn3GZpAYlWy4KU0FMTOZfkyFXyIyYWkKZ4vZWwsZUUWZsNiUbgrf68jppX1U9t+o1ryv12zyOIpzBOVyCBzWowz00oAUMEJ7hFd6cR+fFeXc+lq0FJ585hT9wPn8Aep2MtQ==</latexit>volumefraction<latexit sha1_base64="GA3pj1t8xjD8f0P0H2uFY7Ope5w=">AAAB/3icbVDLSgMxFM3UV62vUcGNm2ARXJUZEXRZdOOygn1AW0omvdOGZpIhyRTL2IW/4saFIm79DXf+jZl2Ftp64MLhnHuTe08Qc6aN5307hZXVtfWN4mZpa3tnd8/dP2homSgKdSq5VK2AaOBMQN0ww6EVKyBRwKEZjG4yvzkGpZkU92YSQzciA8FCRomxUs896hh4MOlY8iQCHCpCM33ac8texZsBLxM/J2WUo9Zzvzp9Se0bwlBOtG77Xmy6KVGGUQ7TUifREBM6IgNoWypIBLqbzvaf4lOr9HEolS1h8Ez9PZGSSOtJFNjOiJihXvQy8T+vnZjwqpsyEScGBJ1/FCYcG4mzMHCfKaCGTywhVDG7K6ZDkoVgIyvZEPzFk5dJ47ziexX/7qJcvc7jKKJjdILOkI8uURXdohqqI4oe0TN6RW/Ok/PivDsf89aCk88coj9wPn8A8diWrg==</latexit><latexit sha1_base64="GA3pj1t8xjD8f0P0H2uFY7Ope5w=">AAAB/3icbVDLSgMxFM3UV62vUcGNm2ARXJUZEXRZdOOygn1AW0omvdOGZpIhyRTL2IW/4saFIm79DXf+jZl2Ftp64MLhnHuTe08Qc6aN5307hZXVtfWN4mZpa3tnd8/dP2homSgKdSq5VK2AaOBMQN0ww6EVKyBRwKEZjG4yvzkGpZkU92YSQzciA8FCRomxUs896hh4MOlY8iQCHCpCM33ac8texZsBLxM/J2WUo9Zzvzp9Se0bwlBOtG77Xmy6KVGGUQ7TUifREBM6IgNoWypIBLqbzvaf4lOr9HEolS1h8Ez9PZGSSOtJFNjOiJihXvQy8T+vnZjwqpsyEScGBJ1/FCYcG4mzMHCfKaCGTywhVDG7K6ZDkoVgIyvZEPzFk5dJ47ziexX/7qJcvc7jKKJjdILOkI8uURXdohqqI4oe0TN6RW/Ok/PivDsf89aCk88coj9wPn8A8diWrg==</latexit><latexit sha1_base64="GA3pj1t8xjD8f0P0H2uFY7Ope5w=">AAAB/3icbVDLSgMxFM3UV62vUcGNm2ARXJUZEXRZdOOygn1AW0omvdOGZpIhyRTL2IW/4saFIm79DXf+jZl2Ftp64MLhnHuTe08Qc6aN5307hZXVtfWN4mZpa3tnd8/dP2homSgKdSq5VK2AaOBMQN0ww6EVKyBRwKEZjG4yvzkGpZkU92YSQzciA8FCRomxUs896hh4MOlY8iQCHCpCM33ac8texZsBLxM/J2WUo9Zzvzp9Se0bwlBOtG77Xmy6KVGGUQ7TUifREBM6IgNoWypIBLqbzvaf4lOr9HEolS1h8Ez9PZGSSOtJFNjOiJihXvQy8T+vnZjwqpsyEScGBJ1/FCYcG4mzMHCfKaCGTywhVDG7K6ZDkoVgIyvZEPzFk5dJ47ziexX/7qJcvc7jKKJjdILOkI8uURXdohqqI4oe0TN6RW/Ok/PivDsf89aCk88coj9wPn8A8diWrg==</latexit><latexit sha1_base64="GA3pj1t8xjD8f0P0H2uFY7Ope5w=">AAAB/3icbVDLSgMxFM3UV62vUcGNm2ARXJUZEXRZdOOygn1AW0omvdOGZpIhyRTL2IW/4saFIm79DXf+jZl2Ftp64MLhnHuTe08Qc6aN5307hZXVtfWN4mZpa3tnd8/dP2homSgKdSq5VK2AaOBMQN0ww6EVKyBRwKEZjG4yvzkGpZkU92YSQzciA8FCRomxUs896hh4MOlY8iQCHCpCM33ac8texZsBLxM/J2WUo9Zzvzp9Se0bwlBOtG77Xmy6KVGGUQ7TUifREBM6IgNoWypIBLqbzvaf4lOr9HEolS1h8Ez9PZGSSOtJFNjOiJihXvQy8T+vnZjwqpsyEScGBJ1/FCYcG4mzMHCfKaCGTywhVDG7K6ZDkoVgIyvZEPzFk5dJ47ziexX/7qJcvc7jKKJjdILOkI8uURXdohqqI4oe0TN6RW/Ok/PivDsf89aCk88coj9wPn8A8diWrg==</latexit>onlyA<latexit sha1_base64="q2kNG6CiQ6iXM8EtQbIh9iy6qh0=">AAAB+HicbVBNSwMxEM36WetHVz16CbaCp7Irgh6rXjxWsB/QLiWbTdvQbLIks2Jd+ku8eFDEqz/Fm//GtN2Dtj4YeLw3w8y8MBHcgOd9Oyura+sbm4Wt4vbO7l7J3T9oGpVqyhpUCaXbITFMcMkawEGwdqIZiUPBWuHoZuq3Hpg2XMl7GCcsiMlA8j6nBKzUc0tdYI+QKSnGuHJVmfTcslf1ZsDLxM9JGeWo99yvbqRoGjMJVBBjOr6XQJARDZwKNil2U8MSQkdkwDqWShIzE2Szwyf4xCoR7ittSwKeqb8nMhIbM45D2xkTGJpFbyr+53VS6F8GGZdJCkzS+aJ+KjAoPE0BR1wzCvbniBOqub0V0yHRhILNqmhD8BdfXibNs6rvVf2783LtOo+jgI7QMTpFPrpANXSL6qiBKErRM3pFb86T8+K8Ox/z1hUnnzlEf+B8/gA0jZLF</latexit><latexit sha1_base64="q2kNG6CiQ6iXM8EtQbIh9iy6qh0=">AAAB+HicbVBNSwMxEM36WetHVz16CbaCp7Irgh6rXjxWsB/QLiWbTdvQbLIks2Jd+ku8eFDEqz/Fm//GtN2Dtj4YeLw3w8y8MBHcgOd9Oyura+sbm4Wt4vbO7l7J3T9oGpVqyhpUCaXbITFMcMkawEGwdqIZiUPBWuHoZuq3Hpg2XMl7GCcsiMlA8j6nBKzUc0tdYI+QKSnGuHJVmfTcslf1ZsDLxM9JGeWo99yvbqRoGjMJVBBjOr6XQJARDZwKNil2U8MSQkdkwDqWShIzE2Szwyf4xCoR7ittSwKeqb8nMhIbM45D2xkTGJpFbyr+53VS6F8GGZdJCkzS+aJ+KjAoPE0BR1wzCvbniBOqub0V0yHRhILNqmhD8BdfXibNs6rvVf2783LtOo+jgI7QMTpFPrpANXSL6qiBKErRM3pFb86T8+K8Ox/z1hUnnzlEf+B8/gA0jZLF</latexit><latexit sha1_base64="q2kNG6CiQ6iXM8EtQbIh9iy6qh0=">AAAB+HicbVBNSwMxEM36WetHVz16CbaCp7Irgh6rXjxWsB/QLiWbTdvQbLIks2Jd+ku8eFDEqz/Fm//GtN2Dtj4YeLw3w8y8MBHcgOd9Oyura+sbm4Wt4vbO7l7J3T9oGpVqyhpUCaXbITFMcMkawEGwdqIZiUPBWuHoZuq3Hpg2XMl7GCcsiMlA8j6nBKzUc0tdYI+QKSnGuHJVmfTcslf1ZsDLxM9JGeWo99yvbqRoGjMJVBBjOr6XQJARDZwKNil2U8MSQkdkwDqWShIzE2Szwyf4xCoR7ittSwKeqb8nMhIbM45D2xkTGJpFbyr+53VS6F8GGZdJCkzS+aJ+KjAoPE0BR1wzCvbniBOqub0V0yHRhILNqmhD8BdfXibNs6rvVf2783LtOo+jgI7QMTpFPrpANXSL6qiBKErRM3pFb86T8+K8Ox/z1hUnnzlEf+B8/gA0jZLF</latexit><latexit sha1_base64="q2kNG6CiQ6iXM8EtQbIh9iy6qh0=">AAAB+HicbVBNSwMxEM36WetHVz16CbaCp7Irgh6rXjxWsB/QLiWbTdvQbLIks2Jd+ku8eFDEqz/Fm//GtN2Dtj4YeLw3w8y8MBHcgOd9Oyura+sbm4Wt4vbO7l7J3T9oGpVqyhpUCaXbITFMcMkawEGwdqIZiUPBWuHoZuq3Hpg2XMl7GCcsiMlA8j6nBKzUc0tdYI+QKSnGuHJVmfTcslf1ZsDLxM9JGeWo99yvbqRoGjMJVBBjOr6XQJARDZwKNil2U8MSQkdkwDqWShIzE2Szwyf4xCoR7ittSwKeqb8nMhIbM45D2xkTGJpFbyr+53VS6F8GGZdJCkzS+aJ+KjAoPE0BR1wzCvbniBOqub0V0yHRhILNqmhD8BdfXibNs6rvVf2783LtOo+jgI7QMTpFPrpANXSL6qiBKErRM3pFb86T8+K8Ox/z1hUnnzlEf+B8/gA0jZLF</latexit>onlyB<latexit sha1_base64="0wfW6byGCwRXraoIg8MOVr9sS2U=">AAAB+HicbVBNSwMxEM36WetHVz16CbaCp7Irgh5LvXisYD+gXUo2m7ah2WRJZsW69Jd48aCIV3+KN/+NabsHbX0w8Hhvhpl5YSK4Ac/7dtbWNza3tgs7xd29/YOSe3jUMirVlDWpEkp3QmKY4JI1gYNgnUQzEoeCtcPxzcxvPzBtuJL3MElYEJOh5ANOCVip75Z6wB4hU1JMcKVemfbdslf15sCrxM9JGeVo9N2vXqRoGjMJVBBjur6XQJARDZwKNi32UsMSQsdkyLqWShIzE2Tzw6f4zCoRHihtSwKeq78nMhIbM4lD2xkTGJllbyb+53VTGFwHGZdJCkzSxaJBKjAoPEsBR1wzCvbniBOqub0V0xHRhILNqmhD8JdfXiWti6rvVf27y3KtnsdRQCfoFJ0jH12hGrpFDdREFKXoGb2iN+fJeXHenY9F65qTzxyjP3A+fwA2E5LG</latexit><latexit sha1_base64="0wfW6byGCwRXraoIg8MOVr9sS2U=">AAAB+HicbVBNSwMxEM36WetHVz16CbaCp7Irgh5LvXisYD+gXUo2m7ah2WRJZsW69Jd48aCIV3+KN/+NabsHbX0w8Hhvhpl5YSK4Ac/7dtbWNza3tgs7xd29/YOSe3jUMirVlDWpEkp3QmKY4JI1gYNgnUQzEoeCtcPxzcxvPzBtuJL3MElYEJOh5ANOCVip75Z6wB4hU1JMcKVemfbdslf15sCrxM9JGeVo9N2vXqRoGjMJVBBjur6XQJARDZwKNi32UsMSQsdkyLqWShIzE2Tzw6f4zCoRHihtSwKeq78nMhIbM4lD2xkTGJllbyb+53VTGFwHGZdJCkzSxaJBKjAoPEsBR1wzCvbniBOqub0V0xHRhILNqmhD8JdfXiWti6rvVf27y3KtnsdRQCfoFJ0jH12hGrpFDdREFKXoGb2iN+fJeXHenY9F65qTzxyjP3A+fwA2E5LG</latexit><latexit sha1_base64="0wfW6byGCwRXraoIg8MOVr9sS2U=">AAAB+HicbVBNSwMxEM36WetHVz16CbaCp7Irgh5LvXisYD+gXUo2m7ah2WRJZsW69Jd48aCIV3+KN/+NabsHbX0w8Hhvhpl5YSK4Ac/7dtbWNza3tgs7xd29/YOSe3jUMirVlDWpEkp3QmKY4JI1gYNgnUQzEoeCtcPxzcxvPzBtuJL3MElYEJOh5ANOCVip75Z6wB4hU1JMcKVemfbdslf15sCrxM9JGeVo9N2vXqRoGjMJVBBjur6XQJARDZwKNi32UsMSQsdkyLqWShIzE2Tzw6f4zCoRHihtSwKeq78nMhIbM4lD2xkTGJllbyb+53VTGFwHGZdJCkzSxaJBKjAoPEsBR1wzCvbniBOqub0V0xHRhILNqmhD8JdfXiWti6rvVf27y3KtnsdRQCfoFJ0jH12hGrpFDdREFKXoGb2iN+fJeXHenY9F65qTzxyjP3A+fwA2E5LG</latexit><latexit sha1_base64="0wfW6byGCwRXraoIg8MOVr9sS2U=">AAAB+HicbVBNSwMxEM36WetHVz16CbaCp7Irgh5LvXisYD+gXUo2m7ah2WRJZsW69Jd48aCIV3+KN/+NabsHbX0w8Hhvhpl5YSK4Ac/7dtbWNza3tgs7xd29/YOSe3jUMirVlDWpEkp3QmKY4JI1gYNgnUQzEoeCtcPxzcxvPzBtuJL3MElYEJOh5ANOCVip75Z6wB4hU1JMcKVemfbdslf15sCrxM9JGeVo9N2vXqRoGjMJVBBjur6XQJARDZwKNi32UsMSQsdkyLqWShIzE2Tzw6f4zCoRHihtSwKeq78nMhIbM4lD2xkTGJllbyb+53VTGFwHGZdJCkzSxaJBKjAoPEsBR1wzCvbniBOqub0V0xHRhILNqmhD8JdfXiWti6rvVf27y3KtnsdRQCfoFJ0jH12hGrpFDdREFKXoGb2iN+fJeXHenY9F65qTzxyjP3A+fwA2E5LG</latexit> inside liquid-like stress granules triggering the aggrega- tion kinetics inside these compartments. The presence of this phase separated compartment can promote the for- mation fibrillar aggregates, and prevent aggregation out- side the stress granules [19, 22]. Thus the corresponding cytotoxic effects of protein aggregates are expected to be strongly localized in space as well. However, it remains elusive whether weak protein interactions are sufficient to cause a significant change in aggregate concentration in the compartment relative to homogeneous aggrega- tion and how the physical parameters of aggregation and phase separation determine the relative enrichment of ag- gregates inside versus outside. Here, we combine knowledge of the kinetics of irre- versible protein aggregation with the theory of liquid- liquid phase separation to develop a model of irreversible assembly of protein fibrils in the presence of droplet- like compartments. We use this model to predict the degree of enrichment of aggregates into the liquid com- partment as a function of the fundamental physical pa- rameters underlying aggregation kinetics and phase sep- aration. We find that relatively weak interactions be- tween the protein monomers and the liquid compartment molecules are sufficient to enrich the concentration of ag- gregates within the liquid compartment by several or- ders of magnitudes relative to homogeneous aggregation (Fig. 1). This strong enrichment of aggregates emerges because the liquid compartment acts as continuous sink of monomers during the aggregation dynamics, thus pro- moting intra-compartment aggregation but suppressing aggregation outside of the compartment. Moreover, we find that aggregate enrichment is more pronounced for larger (smaller) compartments depending on the relative values of the reaction orders for primary and secondary nucleation. Our results suggest that cellular liquid com- partments are ideal to control irreversible protein aggre- gation in space and that the compartment volume, which is determined by the mean concentration of phase sepa- rated protein, represents a relevant control parameter for intra-compartment positioning of aggregates. This con- trol mechanism could be relevant in the context of spatial regulation of other irreversible chemical reactions where liquid compartments act as biomolecular microreactors. MATHEMATICAL MODEL FOR LIQUID COMPARTMENTS CONTROLLING PROTEIN AGGREGATION To capture the interplay between liquid phase sepa- ration and protein aggregation kinetics we start with a model of two coexisting protein phases where monomers partition differently into these phases determined by their relative interactions. We consider the case where the partitioning of monomers is close to equilibrium continu- ously during the kinetics of aggregation. This assumption is well justified since small weakly interacting molecules such as the aggregating monomers diffuse between sec- 2 FIG. 2. Enrichment of monomers and relative de- gree of segregation. (a) The monomer enrichment Γ ((14)) exponentially increases with the relative interaction strength ∆χ (units of kBT ) between the monomers and the A and B molecules which is defined in the SI (section S1). Its char- acteristic increase it set by the degree of phase separation, φI − φII. Enrichment vanishes at the critical point of phase separation (solid line) and increases with the degree of phase separation (dashed line). Enrichment is largest for φI−φII (cid:39) 1 (dash-dotted line). Due to the exponential increase, large en- richment of monomers Γ can already be reached for weak relative interaction energies of a few kBT . (b) The parti- tion degree ξ( ¯φ) = cII m ((2)) describing the concentration fraction of monomers that resides in the minority phase II of the compartment, decreases with the mean volume fraction of A material, ¯φ, along with increasing compartment volume V I( ¯φ). Smaller compartments are thus better in enriching the monomer mass concentration. m/ctot onds and minutes through a cell of size in the order of tens of µm [9, 40], while typical time scales of aggregation in vitro are in the order of hours (see e.g. [41]). Further- more, the diffusion of aggregates is highly hindered as long fibrillar aggregates experience a much larger hydro- dynamic drag force and can get entangled with cytoskele- tal filaments and other assembled fibrils [42, 43]. Finally, at large enough density and size, fibrils start to form solid-like gels [22] further slowing down their mobility. All these effects imply that we may safely neglect diffu- sion of large aggregates and consider the typical case that monomers diffuse quickly relative to their aggregation ki- netics. This separation of time scales suggests that it is natural to first consider the partitioning of monomers into phase separated compartments at equilibrium and then consider small deviations from this equilibrium to understand its consequences for protein aggregation. Phase separation and partitioning of monomers at equilibrium We consider a system of total volume V hosting a sin- gle liquid compartment (a droplet for example) of a con- densed protein phase I of volume V I. The compartment itself forms by liquid-liquid phase separation between the <latexit sha1_base64="vMCmCBvrQqvMkcZe5c78yK19qu0=">AAAB63icbVBNSwMxEJ3Ur1q/qh69BIvgqeyKoMeiF48VbC20S8mm2W5okl2SrFCW/gUvHhTx6h/y5r8x2+5BWx8MPN6bYWZemApurOd9o8ra+sbmVnW7trO7t39QPzzqmiTTlHVoIhLdC4lhgivWsdwK1ks1IzIU7DGc3Bb+4xPThifqwU5TFkgyVjzilNhCGqQxH9YbXtObA68SvyQNKNEe1r8Go4RmkilLBTGm73upDXKiLaeCzWqDzLCU0AkZs76jikhmgnx+6wyfOWWEo0S7UhbP1d8TOZHGTGXoOiWxsVn2CvE/r5/Z6DrIuUozyxRdLIoygW2Ci8fxiGtGrZg6Qqjm7lZMY6IJtS6emgvBX355lXQvmr7X9O8vG62bMo4qnMApnIMPV9CCO2hDByjE8Ayv8IYkekHv6GPRWkHlzDH8Afr8AROIjj8=</latexit><latexit sha1_base64="vMCmCBvrQqvMkcZe5c78yK19qu0=">AAAB63icbVBNSwMxEJ3Ur1q/qh69BIvgqeyKoMeiF48VbC20S8mm2W5okl2SrFCW/gUvHhTx6h/y5r8x2+5BWx8MPN6bYWZemApurOd9o8ra+sbmVnW7trO7t39QPzzqmiTTlHVoIhLdC4lhgivWsdwK1ks1IzIU7DGc3Bb+4xPThifqwU5TFkgyVjzilNhCGqQxH9YbXtObA68SvyQNKNEe1r8Go4RmkilLBTGm73upDXKiLaeCzWqDzLCU0AkZs76jikhmgnx+6wyfOWWEo0S7UhbP1d8TOZHGTGXoOiWxsVn2CvE/r5/Z6DrIuUozyxRdLIoygW2Ci8fxiGtGrZg6Qqjm7lZMY6IJtS6emgvBX355lXQvmr7X9O8vG62bMo4qnMApnIMPV9CCO2hDByjE8Ayv8IYkekHv6GPRWkHlzDH8Afr8AROIjj8=</latexit><latexit sha1_base64="vMCmCBvrQqvMkcZe5c78yK19qu0=">AAAB63icbVBNSwMxEJ3Ur1q/qh69BIvgqeyKoMeiF48VbC20S8mm2W5okl2SrFCW/gUvHhTx6h/y5r8x2+5BWx8MPN6bYWZemApurOd9o8ra+sbmVnW7trO7t39QPzzqmiTTlHVoIhLdC4lhgivWsdwK1ks1IzIU7DGc3Bb+4xPThifqwU5TFkgyVjzilNhCGqQxH9YbXtObA68SvyQNKNEe1r8Go4RmkilLBTGm73upDXKiLaeCzWqDzLCU0AkZs76jikhmgnx+6wyfOWWEo0S7UhbP1d8TOZHGTGXoOiWxsVn2CvE/r5/Z6DrIuUozyxRdLIoygW2Ci8fxiGtGrZg6Qqjm7lZMY6IJtS6emgvBX355lXQvmr7X9O8vG62bMo4qnMApnIMPV9CCO2hDByjE8Ayv8IYkekHv6GPRWkHlzDH8Afr8AROIjj8=</latexit><latexit sha1_base64="vMCmCBvrQqvMkcZe5c78yK19qu0=">AAAB63icbVBNSwMxEJ3Ur1q/qh69BIvgqeyKoMeiF48VbC20S8mm2W5okl2SrFCW/gUvHhTx6h/y5r8x2+5BWx8MPN6bYWZemApurOd9o8ra+sbmVnW7trO7t39QPzzqmiTTlHVoIhLdC4lhgivWsdwK1ks1IzIU7DGc3Bb+4xPThifqwU5TFkgyVjzilNhCGqQxH9YbXtObA68SvyQNKNEe1r8Go4RmkilLBTGm73upDXKiLaeCzWqDzLCU0AkZs76jikhmgnx+6wyfOWWEo0S7UhbP1d8TOZHGTGXoOiWxsVn2CvE/r5/Z6DrIuUozyxRdLIoygW2Ci8fxiGtGrZg6Qqjm7lZMY6IJtS6emgvBX355lXQvmr7X9O8vG62bMo4qnMApnIMPV9CCO2hDByjE8Ayv8IYkekHv6GPRWkHlzDH8Afr8AROIjj8=</latexit>I<latexit sha1_base64="uj1Lu3/A7RVg+FAHZIJDy/GRYGo=">AAAB+HicbVBNSwMxEM3Wr1o/uurRS7AInsquCHosetFbBfsB7VqyabYNTbJLMivWpb/EiwdFvPpTvPlvTNs9aOuDgcd7M8zMCxPBDXjet1NYWV1b3yhulra2d3bL7t5+08SppqxBYxHrdkgME1yxBnAQrJ1oRmQoWCscXU391gPThsfqDsYJCyQZKB5xSsBKPbfcTYb8vgvsEbTMbiY9t+JVvRnwMvFzUkE56j33q9uPaSqZAiqIMR3fSyDIiAZOBZuUuqlhCaEjMmAdSxWRzATZ7PAJPrZKH0extqUAz9TfExmRxoxlaDslgaFZ9Kbif14nhegiyLhKUmCKzhdFqcAQ42kKuM81oyDGlhCqub0V0yHRhILNqmRD8BdfXibN06rvVf3bs0rtMo+jiA7RETpBPjpHNXSN6qiBKErRM3pFb86T8+K8Ox/z1oKTzxygP3A+fwBLqJN9</latexit><latexit sha1_base64="uj1Lu3/A7RVg+FAHZIJDy/GRYGo=">AAAB+HicbVBNSwMxEM3Wr1o/uurRS7AInsquCHosetFbBfsB7VqyabYNTbJLMivWpb/EiwdFvPpTvPlvTNs9aOuDgcd7M8zMCxPBDXjet1NYWV1b3yhulra2d3bL7t5+08SppqxBYxHrdkgME1yxBnAQrJ1oRmQoWCscXU391gPThsfqDsYJCyQZKB5xSsBKPbfcTYb8vgvsEbTMbiY9t+JVvRnwMvFzUkE56j33q9uPaSqZAiqIMR3fSyDIiAZOBZuUuqlhCaEjMmAdSxWRzATZ7PAJPrZKH0extqUAz9TfExmRxoxlaDslgaFZ9Kbif14nhegiyLhKUmCKzhdFqcAQ42kKuM81oyDGlhCqub0V0yHRhILNqmRD8BdfXibN06rvVf3bs0rtMo+jiA7RETpBPjpHNXSN6qiBKErRM3pFb86T8+K8Ox/z1oKTzxygP3A+fwBLqJN9</latexit><latexit sha1_base64="uj1Lu3/A7RVg+FAHZIJDy/GRYGo=">AAAB+HicbVBNSwMxEM3Wr1o/uurRS7AInsquCHosetFbBfsB7VqyabYNTbJLMivWpb/EiwdFvPpTvPlvTNs9aOuDgcd7M8zMCxPBDXjet1NYWV1b3yhulra2d3bL7t5+08SppqxBYxHrdkgME1yxBnAQrJ1oRmQoWCscXU391gPThsfqDsYJCyQZKB5xSsBKPbfcTYb8vgvsEbTMbiY9t+JVvRnwMvFzUkE56j33q9uPaSqZAiqIMR3fSyDIiAZOBZuUuqlhCaEjMmAdSxWRzATZ7PAJPrZKH0extqUAz9TfExmRxoxlaDslgaFZ9Kbif14nhegiyLhKUmCKzhdFqcAQ42kKuM81oyDGlhCqub0V0yHRhILNqmRD8BdfXibN06rvVf3bs0rtMo+jiA7RETpBPjpHNXSN6qiBKErRM3pFb86T8+K8Ox/z1oKTzxygP3A+fwBLqJN9</latexit><latexit sha1_base64="uj1Lu3/A7RVg+FAHZIJDy/GRYGo=">AAAB+HicbVBNSwMxEM3Wr1o/uurRS7AInsquCHosetFbBfsB7VqyabYNTbJLMivWpb/EiwdFvPpTvPlvTNs9aOuDgcd7M8zMCxPBDXjet1NYWV1b3yhulra2d3bL7t5+08SppqxBYxHrdkgME1yxBnAQrJ1oRmQoWCscXU391gPThsfqDsYJCyQZKB5xSsBKPbfcTYb8vgvsEbTMbiY9t+JVvRnwMvFzUkE56j33q9uPaSqZAiqIMR3fSyDIiAZOBZuUuqlhCaEjMmAdSxWRzATZ7PAJPrZKH0extqUAz9TfExmRxoxlaDslgaFZ9Kbif14nhegiyLhKUmCKzhdFqcAQ42kKuM81oyDGlhCqub0V0yHRhILNqmRD8BdfXibN06rvVf3bs0rtMo+jiA7RETpBPjpHNXSN6qiBKErRM3pFb86T8+K8Ox/z1oKTzxygP3A+fwBLqJN9</latexit>II<latexit sha1_base64="Q+wlYgKtwjfDzs+1/njkyYyv+wU=">AAAB+XicbVBNSwMxEM3Wr1q/Vj16CRbBU9kVQY9FL/ZWwX5Au5Zsmm1Dk+ySzBbL0n/ixYMiXv0n3vw3pu0etPXBwOO9GWbmhYngBjzv2ymsrW9sbhW3Szu7e/sH7uFR08SppqxBYxHrdkgME1yxBnAQrJ1oRmQoWCsc3c781phpw2P1AJOEBZIMFI84JWClnut2kyF/7AJ7Ai2zWm3ac8texZsDrxI/J2WUo95zv7r9mKaSKaCCGNPxvQSCjGjgVLBpqZsalhA6IgPWsVQRyUyQzS+f4jOr9HEUa1sK8Fz9PZERacxEhrZTEhiaZW8m/ud1Uoiug4yrJAWm6GJRlAoMMZ7FgPtcMwpiYgmhmttbMR0STSjYsEo2BH/55VXSvKj4XsW/vyxXb/I4iugEnaJz5KMrVEV3qI4aiKIxekav6M3JnBfn3flYtBacfOYY/YHz+QPlc5PQ</latexit><latexit sha1_base64="Q+wlYgKtwjfDzs+1/njkyYyv+wU=">AAAB+XicbVBNSwMxEM3Wr1q/Vj16CRbBU9kVQY9FL/ZWwX5Au5Zsmm1Dk+ySzBbL0n/ixYMiXv0n3vw3pu0etPXBwOO9GWbmhYngBjzv2ymsrW9sbhW3Szu7e/sH7uFR08SppqxBYxHrdkgME1yxBnAQrJ1oRmQoWCsc3c781phpw2P1AJOEBZIMFI84JWClnut2kyF/7AJ7Ai2zWm3ac8texZsDrxI/J2WUo95zv7r9mKaSKaCCGNPxvQSCjGjgVLBpqZsalhA6IgPWsVQRyUyQzS+f4jOr9HEUa1sK8Fz9PZERacxEhrZTEhiaZW8m/ud1Uoiug4yrJAWm6GJRlAoMMZ7FgPtcMwpiYgmhmttbMR0STSjYsEo2BH/55VXSvKj4XsW/vyxXb/I4iugEnaJz5KMrVEV3qI4aiKIxekav6M3JnBfn3flYtBacfOYY/YHz+QPlc5PQ</latexit><latexit sha1_base64="Q+wlYgKtwjfDzs+1/njkyYyv+wU=">AAAB+XicbVBNSwMxEM3Wr1q/Vj16CRbBU9kVQY9FL/ZWwX5Au5Zsmm1Dk+ySzBbL0n/ixYMiXv0n3vw3pu0etPXBwOO9GWbmhYngBjzv2ymsrW9sbhW3Szu7e/sH7uFR08SppqxBYxHrdkgME1yxBnAQrJ1oRmQoWCsc3c781phpw2P1AJOEBZIMFI84JWClnut2kyF/7AJ7Ai2zWm3ac8texZsDrxI/J2WUo95zv7r9mKaSKaCCGNPxvQSCjGjgVLBpqZsalhA6IgPWsVQRyUyQzS+f4jOr9HEUa1sK8Fz9PZERacxEhrZTEhiaZW8m/ud1Uoiug4yrJAWm6GJRlAoMMZ7FgPtcMwpiYgmhmttbMR0STSjYsEo2BH/55VXSvKj4XsW/vyxXb/I4iugEnaJz5KMrVEV3qI4aiKIxekav6M3JnBfn3flYtBacfOYY/YHz+QPlc5PQ</latexit><latexit sha1_base64="Q+wlYgKtwjfDzs+1/njkyYyv+wU=">AAAB+XicbVBNSwMxEM3Wr1q/Vj16CRbBU9kVQY9FL/ZWwX5Au5Zsmm1Dk+ySzBbL0n/ixYMiXv0n3vw3pu0etPXBwOO9GWbmhYngBjzv2ymsrW9sbhW3Szu7e/sH7uFR08SppqxBYxHrdkgME1yxBnAQrJ1oRmQoWCsc3c781phpw2P1AJOEBZIMFI84JWClnut2kyF/7AJ7Ai2zWm3ac8texZsDrxI/J2WUo95zv7r9mKaSKaCCGNPxvQSCjGjgVLBpqZsalhA6IgPWsVQRyUyQzS+f4jOr9HEUa1sK8Fz9PZERacxEhrZTEhiaZW8m/ud1Uoiug4yrJAWm6GJRlAoMMZ7FgPtcMwpiYgmhmttbMR0STSjYsEo2BH/55VXSvKj4XsW/vyxXb/I4iugEnaJz5KMrVEV3qI4aiKIxekav6M3JnBfn3flYtBacfOYY/YHz+QPlc5PQ</latexit>0<latexit sha1_base64="6ZTwbptvK00HUiMuNssEoeJJPkc=">AAAB6HicbVBNS8NAEJ3Ur1q/qh69LBbBU0lEqMeiF48t2A9oQ9lsJ+3azSbsboQS+gu8eFDEqz/Jm//GbZuDtj4YeLw3w8y8IBFcG9f9dgobm1vbO8Xd0t7+weFR+fikreNUMWyxWMSqG1CNgktsGW4EdhOFNAoEdoLJ3dzvPKHSPJYPZpqgH9GR5CFn1Fip6Q7KFbfqLkDWiZeTCuRoDMpf/WHM0gilYYJq3fPcxPgZVYYzgbNSP9WYUDahI+xZKmmE2s8Wh87IhVWGJIyVLWnIQv09kdFI62kU2M6ImrFe9ebif14vNeGNn3GZpAYlWy4KU0FMTOZfkyFXyIyYWkKZ4vZWwsZUUWZsNiUbgrf68jppX1U9t+o1ryv12zyOIpzBOVyCBzWowz00oAUMEJ7hFd6cR+fFeXc+lq0FJ585hT9wPn8AeRmMtA==</latexit><latexit sha1_base64="6ZTwbptvK00HUiMuNssEoeJJPkc=">AAAB6HicbVBNS8NAEJ3Ur1q/qh69LBbBU0lEqMeiF48t2A9oQ9lsJ+3azSbsboQS+gu8eFDEqz/Jm//GbZuDtj4YeLw3w8y8IBFcG9f9dgobm1vbO8Xd0t7+weFR+fikreNUMWyxWMSqG1CNgktsGW4EdhOFNAoEdoLJ3dzvPKHSPJYPZpqgH9GR5CFn1Fip6Q7KFbfqLkDWiZeTCuRoDMpf/WHM0gilYYJq3fPcxPgZVYYzgbNSP9WYUDahI+xZKmmE2s8Wh87IhVWGJIyVLWnIQv09kdFI62kU2M6ImrFe9ebif14vNeGNn3GZpAYlWy4KU0FMTOZfkyFXyIyYWkKZ4vZWwsZUUWZsNiUbgrf68jppX1U9t+o1ryv12zyOIpzBOVyCBzWowz00oAUMEJ7hFd6cR+fFeXc+lq0FJ585hT9wPn8AeRmMtA==</latexit><latexit sha1_base64="6ZTwbptvK00HUiMuNssEoeJJPkc=">AAAB6HicbVBNS8NAEJ3Ur1q/qh69LBbBU0lEqMeiF48t2A9oQ9lsJ+3azSbsboQS+gu8eFDEqz/Jm//GbZuDtj4YeLw3w8y8IBFcG9f9dgobm1vbO8Xd0t7+weFR+fikreNUMWyxWMSqG1CNgktsGW4EdhOFNAoEdoLJ3dzvPKHSPJYPZpqgH9GR5CFn1Fip6Q7KFbfqLkDWiZeTCuRoDMpf/WHM0gilYYJq3fPcxPgZVYYzgbNSP9WYUDahI+xZKmmE2s8Wh87IhVWGJIyVLWnIQv09kdFI62kU2M6ImrFe9ebif14vNeGNn3GZpAYlWy4KU0FMTOZfkyFXyIyYWkKZ4vZWwsZUUWZsNiUbgrf68jppX1U9t+o1ryv12zyOIpzBOVyCBzWowz00oAUMEJ7hFd6cR+fFeXc+lq0FJ585hT9wPn8AeRmMtA==</latexit><latexit sha1_base64="6ZTwbptvK00HUiMuNssEoeJJPkc=">AAAB6HicbVBNS8NAEJ3Ur1q/qh69LBbBU0lEqMeiF48t2A9oQ9lsJ+3azSbsboQS+gu8eFDEqz/Jm//GbZuDtj4YeLw3w8y8IBFcG9f9dgobm1vbO8Xd0t7+weFR+fikreNUMWyxWMSqG1CNgktsGW4EdhOFNAoEdoLJ3dzvPKHSPJYPZpqgH9GR5CFn1Fip6Q7KFbfqLkDWiZeTCuRoDMpf/WHM0gilYYJq3fPcxPgZVYYzgbNSP9WYUDahI+xZKmmE2s8Wh87IhVWGJIyVLWnIQv09kdFI62kU2M6ImrFe9ebif14vNeGNn3GZpAYlWy4KU0FMTOZfkyFXyIyYWkKZ4vZWwsZUUWZsNiUbgrf68jppX1U9t+o1ryv12zyOIpzBOVyCBzWowz00oAUMEJ7hFd6cR+fFeXc+lq0FJ585hT9wPn8AeRmMtA==</latexit>1<latexit sha1_base64="l9eImvYcFOKpzEDji/n9jPDeWb8=">AAAB6HicbVBNS8NAEJ3Ur1q/qh69LBbBU0lEqMeiF48t2A9oQ9lsJ+3azSbsboQS+gu8eFDEqz/Jm//GbZuDtj4YeLw3w8y8IBFcG9f9dgobm1vbO8Xd0t7+weFR+fikreNUMWyxWMSqG1CNgktsGW4EdhOFNAoEdoLJ3dzvPKHSPJYPZpqgH9GR5CFn1Fip6Q3KFbfqLkDWiZeTCuRoDMpf/WHM0gilYYJq3fPcxPgZVYYzgbNSP9WYUDahI+xZKmmE2s8Wh87IhVWGJIyVLWnIQv09kdFI62kU2M6ImrFe9ebif14vNeGNn3GZpAYlWy4KU0FMTOZfkyFXyIyYWkKZ4vZWwsZUUWZsNiUbgrf68jppX1U9t+o1ryv12zyOIpzBOVyCBzWowz00oAUMEJ7hFd6cR+fFeXc+lq0FJ585hT9wPn8Aep2MtQ==</latexit><latexit sha1_base64="l9eImvYcFOKpzEDji/n9jPDeWb8=">AAAB6HicbVBNS8NAEJ3Ur1q/qh69LBbBU0lEqMeiF48t2A9oQ9lsJ+3azSbsboQS+gu8eFDEqz/Jm//GbZuDtj4YeLw3w8y8IBFcG9f9dgobm1vbO8Xd0t7+weFR+fikreNUMWyxWMSqG1CNgktsGW4EdhOFNAoEdoLJ3dzvPKHSPJYPZpqgH9GR5CFn1Fip6Q3KFbfqLkDWiZeTCuRoDMpf/WHM0gilYYJq3fPcxPgZVYYzgbNSP9WYUDahI+xZKmmE2s8Wh87IhVWGJIyVLWnIQv09kdFI62kU2M6ImrFe9ebif14vNeGNn3GZpAYlWy4KU0FMTOZfkyFXyIyYWkKZ4vZWwsZUUWZsNiUbgrf68jppX1U9t+o1ryv12zyOIpzBOVyCBzWowz00oAUMEJ7hFd6cR+fFeXc+lq0FJ585hT9wPn8Aep2MtQ==</latexit><latexit sha1_base64="l9eImvYcFOKpzEDji/n9jPDeWb8=">AAAB6HicbVBNS8NAEJ3Ur1q/qh69LBbBU0lEqMeiF48t2A9oQ9lsJ+3azSbsboQS+gu8eFDEqz/Jm//GbZuDtj4YeLw3w8y8IBFcG9f9dgobm1vbO8Xd0t7+weFR+fikreNUMWyxWMSqG1CNgktsGW4EdhOFNAoEdoLJ3dzvPKHSPJYPZpqgH9GR5CFn1Fip6Q3KFbfqLkDWiZeTCuRoDMpf/WHM0gilYYJq3fPcxPgZVYYzgbNSP9WYUDahI+xZKmmE2s8Wh87IhVWGJIyVLWnIQv09kdFI62kU2M6ImrFe9ebif14vNeGNn3GZpAYlWy4KU0FMTOZfkyFXyIyYWkKZ4vZWwsZUUWZsNiUbgrf68jppX1U9t+o1ryv12zyOIpzBOVyCBzWowz00oAUMEJ7hFd6cR+fFeXc+lq0FJ585hT9wPn8Aep2MtQ==</latexit><latexit sha1_base64="l9eImvYcFOKpzEDji/n9jPDeWb8=">AAAB6HicbVBNS8NAEJ3Ur1q/qh69LBbBU0lEqMeiF48t2A9oQ9lsJ+3azSbsboQS+gu8eFDEqz/Jm//GbZuDtj4YeLw3w8y8IBFcG9f9dgobm1vbO8Xd0t7+weFR+fikreNUMWyxWMSqG1CNgktsGW4EdhOFNAoEdoLJ3dzvPKHSPJYPZpqgH9GR5CFn1Fip6Q3KFbfqLkDWiZeTCuRoDMpf/WHM0gilYYJq3fPcxPgZVYYzgbNSP9WYUDahI+xZKmmE2s8Wh87IhVWGJIyVLWnIQv09kdFI62kU2M6ImrFe9ebif14vNeGNn3GZpAYlWy4KU0FMTOZfkyFXyIyYWkKZ4vZWwsZUUWZsNiUbgrf68jppX1U9t+o1ryv12zyOIpzBOVyCBzWowz00oAUMEJ7hFd6cR+fFeXc+lq0FJ585hT9wPn8Aep2MtQ==</latexit>volumefraction<latexit sha1_base64="GA3pj1t8xjD8f0P0H2uFY7Ope5w=">AAAB/3icbVDLSgMxFM3UV62vUcGNm2ARXJUZEXRZdOOygn1AW0omvdOGZpIhyRTL2IW/4saFIm79DXf+jZl2Ftp64MLhnHuTe08Qc6aN5307hZXVtfWN4mZpa3tnd8/dP2homSgKdSq5VK2AaOBMQN0ww6EVKyBRwKEZjG4yvzkGpZkU92YSQzciA8FCRomxUs896hh4MOlY8iQCHCpCM33ac8texZsBLxM/J2WUo9Zzvzp9Se0bwlBOtG77Xmy6KVGGUQ7TUifREBM6IgNoWypIBLqbzvaf4lOr9HEolS1h8Ez9PZGSSOtJFNjOiJihXvQy8T+vnZjwqpsyEScGBJ1/FCYcG4mzMHCfKaCGTywhVDG7K6ZDkoVgIyvZEPzFk5dJ47ziexX/7qJcvc7jKKJjdILOkI8uURXdohqqI4oe0TN6RW/Ok/PivDsf89aCk88coj9wPn8A8diWrg==</latexit><latexit sha1_base64="GA3pj1t8xjD8f0P0H2uFY7Ope5w=">AAAB/3icbVDLSgMxFM3UV62vUcGNm2ARXJUZEXRZdOOygn1AW0omvdOGZpIhyRTL2IW/4saFIm79DXf+jZl2Ftp64MLhnHuTe08Qc6aN5307hZXVtfWN4mZpa3tnd8/dP2homSgKdSq5VK2AaOBMQN0ww6EVKyBRwKEZjG4yvzkGpZkU92YSQzciA8FCRomxUs896hh4MOlY8iQCHCpCM33ac8texZsBLxM/J2WUo9Zzvzp9Se0bwlBOtG77Xmy6KVGGUQ7TUifREBM6IgNoWypIBLqbzvaf4lOr9HEolS1h8Ez9PZGSSOtJFNjOiJihXvQy8T+vnZjwqpsyEScGBJ1/FCYcG4mzMHCfKaCGTywhVDG7K6ZDkoVgIyvZEPzFk5dJ47ziexX/7qJcvc7jKKJjdILOkI8uURXdohqqI4oe0TN6RW/Ok/PivDsf89aCk88coj9wPn8A8diWrg==</latexit><latexit sha1_base64="GA3pj1t8xjD8f0P0H2uFY7Ope5w=">AAAB/3icbVDLSgMxFM3UV62vUcGNm2ARXJUZEXRZdOOygn1AW0omvdOGZpIhyRTL2IW/4saFIm79DXf+jZl2Ftp64MLhnHuTe08Qc6aN5307hZXVtfWN4mZpa3tnd8/dP2homSgKdSq5VK2AaOBMQN0ww6EVKyBRwKEZjG4yvzkGpZkU92YSQzciA8FCRomxUs896hh4MOlY8iQCHCpCM33ac8texZsBLxM/J2WUo9Zzvzp9Se0bwlBOtG77Xmy6KVGGUQ7TUifREBM6IgNoWypIBLqbzvaf4lOr9HEolS1h8Ez9PZGSSOtJFNjOiJihXvQy8T+vnZjwqpsyEScGBJ1/FCYcG4mzMHCfKaCGTywhVDG7K6ZDkoVgIyvZEPzFk5dJ47ziexX/7qJcvc7jKKJjdILOkI8uURXdohqqI4oe0TN6RW/Ok/PivDsf89aCk88coj9wPn8A8diWrg==</latexit><latexit sha1_base64="GA3pj1t8xjD8f0P0H2uFY7Ope5w=">AAAB/3icbVDLSgMxFM3UV62vUcGNm2ARXJUZEXRZdOOygn1AW0omvdOGZpIhyRTL2IW/4saFIm79DXf+jZl2Ftp64MLhnHuTe08Qc6aN5307hZXVtfWN4mZpa3tnd8/dP2homSgKdSq5VK2AaOBMQN0ww6EVKyBRwKEZjG4yvzkGpZkU92YSQzciA8FCRomxUs896hh4MOlY8iQCHCpCM33ac8texZsBLxM/J2WUo9Zzvzp9Se0bwlBOtG77Xmy6KVGGUQ7TUifREBM6IgNoWypIBLqbzvaf4lOr9HEolS1h8Ez9PZGSSOtJFNjOiJihXvQy8T+vnZjwqpsyEScGBJ1/FCYcG4mzMHCfKaCGTywhVDG7K6ZDkoVgIyvZEPzFk5dJ47ziexX/7qJcvc7jKKJjdILOkI8uURXdohqqI4oe0TN6RW/Ok/PivDsf89aCk88coj9wPn8A8diWrg==</latexit>onlyA<latexit sha1_base64="q2kNG6CiQ6iXM8EtQbIh9iy6qh0=">AAAB+HicbVBNSwMxEM36WetHVz16CbaCp7Irgh6rXjxWsB/QLiWbTdvQbLIks2Jd+ku8eFDEqz/Fm//GtN2Dtj4YeLw3w8y8MBHcgOd9Oyura+sbm4Wt4vbO7l7J3T9oGpVqyhpUCaXbITFMcMkawEGwdqIZiUPBWuHoZuq3Hpg2XMl7GCcsiMlA8j6nBKzUc0tdYI+QKSnGuHJVmfTcslf1ZsDLxM9JGeWo99yvbqRoGjMJVBBjOr6XQJARDZwKNil2U8MSQkdkwDqWShIzE2Szwyf4xCoR7ittSwKeqb8nMhIbM45D2xkTGJpFbyr+53VS6F8GGZdJCkzS+aJ+KjAoPE0BR1wzCvbniBOqub0V0yHRhILNqmhD8BdfXibNs6rvVf2783LtOo+jgI7QMTpFPrpANXSL6qiBKErRM3pFb86T8+K8Ox/z1hUnnzlEf+B8/gA0jZLF</latexit><latexit sha1_base64="q2kNG6CiQ6iXM8EtQbIh9iy6qh0=">AAAB+HicbVBNSwMxEM36WetHVz16CbaCp7Irgh6rXjxWsB/QLiWbTdvQbLIks2Jd+ku8eFDEqz/Fm//GtN2Dtj4YeLw3w8y8MBHcgOd9Oyura+sbm4Wt4vbO7l7J3T9oGpVqyhpUCaXbITFMcMkawEGwdqIZiUPBWuHoZuq3Hpg2XMl7GCcsiMlA8j6nBKzUc0tdYI+QKSnGuHJVmfTcslf1ZsDLxM9JGeWo99yvbqRoGjMJVBBjOr6XQJARDZwKNil2U8MSQkdkwDqWShIzE2Szwyf4xCoR7ittSwKeqb8nMhIbM45D2xkTGJpFbyr+53VS6F8GGZdJCkzS+aJ+KjAoPE0BR1wzCvbniBOqub0V0yHRhILNqmhD8BdfXibNs6rvVf2783LtOo+jgI7QMTpFPrpANXSL6qiBKErRM3pFb86T8+K8Ox/z1hUnnzlEf+B8/gA0jZLF</latexit><latexit sha1_base64="q2kNG6CiQ6iXM8EtQbIh9iy6qh0=">AAAB+HicbVBNSwMxEM36WetHVz16CbaCp7Irgh6rXjxWsB/QLiWbTdvQbLIks2Jd+ku8eFDEqz/Fm//GtN2Dtj4YeLw3w8y8MBHcgOd9Oyura+sbm4Wt4vbO7l7J3T9oGpVqyhpUCaXbITFMcMkawEGwdqIZiUPBWuHoZuq3Hpg2XMl7GCcsiMlA8j6nBKzUc0tdYI+QKSnGuHJVmfTcslf1ZsDLxM9JGeWo99yvbqRoGjMJVBBjOr6XQJARDZwKNil2U8MSQkdkwDqWShIzE2Szwyf4xCoR7ittSwKeqb8nMhIbM45D2xkTGJpFbyr+53VS6F8GGZdJCkzS+aJ+KjAoPE0BR1wzCvbniBOqub0V0yHRhILNqmhD8BdfXibNs6rvVf2783LtOo+jgI7QMTpFPrpANXSL6qiBKErRM3pFb86T8+K8Ox/z1hUnnzlEf+B8/gA0jZLF</latexit><latexit sha1_base64="q2kNG6CiQ6iXM8EtQbIh9iy6qh0=">AAAB+HicbVBNSwMxEM36WetHVz16CbaCp7Irgh6rXjxWsB/QLiWbTdvQbLIks2Jd+ku8eFDEqz/Fm//GtN2Dtj4YeLw3w8y8MBHcgOd9Oyura+sbm4Wt4vbO7l7J3T9oGpVqyhpUCaXbITFMcMkawEGwdqIZiUPBWuHoZuq3Hpg2XMl7GCcsiMlA8j6nBKzUc0tdYI+QKSnGuHJVmfTcslf1ZsDLxM9JGeWo99yvbqRoGjMJVBBjOr6XQJARDZwKNil2U8MSQkdkwDqWShIzE2Szwyf4xCoR7ittSwKeqb8nMhIbM45D2xkTGJpFbyr+53VS6F8GGZdJCkzS+aJ+KjAoPE0BR1wzCvbniBOqub0V0yHRhILNqmhD8BdfXibNs6rvVf2783LtOo+jgI7QMTpFPrpANXSL6qiBKErRM3pFb86T8+K8Ox/z1hUnnzlEf+B8/gA0jZLF</latexit>onlyB<latexit sha1_base64="0wfW6byGCwRXraoIg8MOVr9sS2U=">AAAB+HicbVBNSwMxEM36WetHVz16CbaCp7Irgh5LvXisYD+gXUo2m7ah2WRJZsW69Jd48aCIV3+KN/+NabsHbX0w8Hhvhpl5YSK4Ac/7dtbWNza3tgs7xd29/YOSe3jUMirVlDWpEkp3QmKY4JI1gYNgnUQzEoeCtcPxzcxvPzBtuJL3MElYEJOh5ANOCVip75Z6wB4hU1JMcKVemfbdslf15sCrxM9JGeVo9N2vXqRoGjMJVBBjur6XQJARDZwKNi32UsMSQsdkyLqWShIzE2Tzw6f4zCoRHihtSwKeq78nMhIbM4lD2xkTGJllbyb+53VTGFwHGZdJCkzSxaJBKjAoPEsBR1wzCvbniBOqub0V0xHRhILNqmhD8JdfXiWti6rvVf27y3KtnsdRQCfoFJ0jH12hGrpFDdREFKXoGb2iN+fJeXHenY9F65qTzxyjP3A+fwA2E5LG</latexit><latexit sha1_base64="0wfW6byGCwRXraoIg8MOVr9sS2U=">AAAB+HicbVBNSwMxEM36WetHVz16CbaCp7Irgh5LvXisYD+gXUo2m7ah2WRJZsW69Jd48aCIV3+KN/+NabsHbX0w8Hhvhpl5YSK4Ac/7dtbWNza3tgs7xd29/YOSe3jUMirVlDWpEkp3QmKY4JI1gYNgnUQzEoeCtcPxzcxvPzBtuJL3MElYEJOh5ANOCVip75Z6wB4hU1JMcKVemfbdslf15sCrxM9JGeVo9N2vXqRoGjMJVBBjur6XQJARDZwKNi32UsMSQsdkyLqWShIzE2Tzw6f4zCoRHihtSwKeq78nMhIbM4lD2xkTGJllbyb+53VTGFwHGZdJCkzSxaJBKjAoPEsBR1wzCvbniBOqub0V0xHRhILNqmhD8JdfXiWti6rvVf27y3KtnsdRQCfoFJ0jH12hGrpFDdREFKXoGb2iN+fJeXHenY9F65qTzxyjP3A+fwA2E5LG</latexit><latexit sha1_base64="0wfW6byGCwRXraoIg8MOVr9sS2U=">AAAB+HicbVBNSwMxEM36WetHVz16CbaCp7Irgh5LvXisYD+gXUo2m7ah2WRJZsW69Jd48aCIV3+KN/+NabsHbX0w8Hhvhpl5YSK4Ac/7dtbWNza3tgs7xd29/YOSe3jUMirVlDWpEkp3QmKY4JI1gYNgnUQzEoeCtcPxzcxvPzBtuJL3MElYEJOh5ANOCVip75Z6wB4hU1JMcKVemfbdslf15sCrxM9JGeVo9N2vXqRoGjMJVBBjur6XQJARDZwKNi32UsMSQsdkyLqWShIzE2Tzw6f4zCoRHihtSwKeq78nMhIbM4lD2xkTGJllbyb+53VTGFwHGZdJCkzSxaJBKjAoPEsBR1wzCvbniBOqub0V0xHRhILNqmhD8JdfXiWti6rvVf27y3KtnsdRQCfoFJ0jH12hGrpFDdREFKXoGb2iN+fJeXHenY9F65qTzxyjP3A+fwA2E5LG</latexit><latexit sha1_base64="0wfW6byGCwRXraoIg8MOVr9sS2U=">AAAB+HicbVBNSwMxEM36WetHVz16CbaCp7Irgh5LvXisYD+gXUo2m7ah2WRJZsW69Jd48aCIV3+KN/+NabsHbX0w8Hhvhpl5YSK4Ac/7dtbWNza3tgs7xd29/YOSe3jUMirVlDWpEkp3QmKY4JI1gYNgnUQzEoeCtcPxzcxvPzBtuJL3MElYEJOh5ANOCVip75Z6wB4hU1JMcKVemfbdslf15sCrxM9JGeVo9N2vXqRoGjMJVBBjur6XQJARDZwKNi32UsMSQsdkyLqWShIzE2Tzw6f4zCoRHihtSwKeq78nMhIbM4lD2xkTGJllbyb+53VTGFwHGZdJCkzSxaJBKjAoPEsBR1wzCvbniBOqub0V0xHRhILNqmhD8JdfXiWti6rvVf27y3KtnsdRQCfoFJ0jH12hGrpFDdREFKXoGb2iN+fJeXHenY9F65qTzxyjP3A+fwA2E5LG</latexit>monomerenrichmentrelativeinteraction012345101234(a)partitiondegree⇠(¯)volumefraction¯0100.51(b)1/III two components A and B. Compartment I is composed mostly of the protein component A and a small fraction of protein component B, while compartment II has a small amount of protein A and a large amount of protein B, as depicted in Fig. 1. For simplicity, we discuss the case of an incompressible system where the aggregating monomers 'm' and aggre- gates 'a' are dilute, i.e., cmνm (cid:28) 1 and caνa (cid:28) 1, with cm and ca denoting the concentrations of monomers and aggregates and νm and νa are the respective molecular volumes. For instance, typical values of volume fractions for monomers of Amyloid-β, cmνm (radius of gyration in the range 1 -- 2 nm [44]), at physiological concentra- tions between 100pM to 1nM are in the range of 10−9 to 10−8. The assumption of dilute monomers and ag- gregates implies that for an incompressible system the volume fractions φA and φB of the protein components A and B obey φA + φB = 1 − cmνm − caνa (cid:39) 1, where we abbreviate φA = φ in the following. The monomers may partition differently into the respective minority and majority phases, but, due to their dilute concentrations, they do not affect the degree of phase separation. Un- der these circumstances, the partitioning of monomers in the two phases is solely governed by the relative inter- action strength ∆χ between the monomers with the A and the B components, respectively. If ∆χ is large and positive, monomers favor the presence of the majority component A in compartment I. In this case we expect a more pronounced partitioning of monomers into com- partment I. Contrariwise, when ∆χ is large and negative, monomers favorably partition into compartment II. The degree of monomer enrichment at equilibrium can be cal- culated using the condition that the chemical potentials of monomers associated with compartment I and II are balanced (see Fig. 2(a), SI, section S1, for the derivation), and allows us to define the monomer enrichment factor (cid:104) νm ν ∆χ(cid:0)φI − φII(cid:1)(cid:105) Γ ≡ cI m m (cid:39) exp cII , (1) m, cI m = (cid:0)cI where cI m are the monomer concentrations in phases I and II, respectively, ν denotes the molecular volume of A and B molecules, and φI − φII ∈ [0, 1] is the de- gree of phase separation of the A-component. Then the relative partitioning of the total monomer concentration, ctot m = ξ( ¯φ)Γctot cI degree mV II(cid:1) /V , is given by the expressions ξ(cid:0) ¯φ(cid:1) = 1 + (Γ − 1)V I(cid:0) ¯φ(cid:1) /V m , where the partition m = ξ( ¯φ)ctot m and cII mV I + cII (2) 1 captures the impact of the relative size of the compart- ment volume V I( ¯φ)/V . The volume of the compartment I is in turn controlled by the mean volume fraction ¯φ of A molecules in the system in terms of the relationship V I( ¯φ) = V ( ¯φ−φII)/(φI−φII). The relationships (14) and (2) characterize the relatively rapid physics of monomer partitioning at equilibrium and serve as the background 3 atop which we consider out-of-equilibrium aggregation kinetics. Model for protein aggregation kinetics coupled to non-equilibrium monomer partitioning Due to the separation of time scales of monomer dif- fusion and monomer aggregation, the partitioning of monomers into the compartment is close to equilibrium at all times of the aggregation kinetics and thus the rela- tive fraction of monomers is approximately governed by monomer enrichment Γ, (14). However, as the aggre- gation kinetics decreases the amount of monomers inside each phase, aggregation couples to the partitioning which cause the generation of diffusive fluxes J α that attempt to maintain the monomer partitioning close to equilib- rium. In the limit of a sharp interface separating the liquid compartment from the bulk, so that there is no aggregation at the interface, J I = −J II. Furthermore, to linear order, the flux J α between the phases is pro- portional to the difference of monomer partitioning with respect the equilibrium value Γ (see SI, section S2, for the derivation) and is of the form: = −kα (cid:0)M I J α V α(R) (cid:1) , m − ΓM II m (3a) m = cα where M α mmm (with mm as monomer mass) is the monomer mass concentration in compartment α = I, II, and kα denotes the rate at which monomer partitioning relaxes back to the equilibrium given by (14). In each phase, irreversible protein aggregation of fib- rillar structures occurs as a consequence of both primary and secondary nucleation, and growth of aggregates via their ends, each event occurring with a rate k1, k2, and k+ [5, 45 -- 47]. We see that the key term in our model is the difference between the monomer concentration in- side and outside of the compartment which leads to the diffusive flux of monomers J α between the phases ((3a)), which connects the effects of phase separation and pro- tein aggregation. The coupled equations describing the protein aggregation in both phases can be written as dcα a (t) dt dM α a (t) dt dM α m(t) dt = k1 M α m(t)n1 + k2 M α m(t)n2 M α a (t) , = 2k+ M α m(t) cα a (t) , = −2k+ M α m(t) cα a (t) + J α V α . (3b) (3c) (3d) Here, (3b) describes the formation of new fibrils in each compartment through primary nucleation, fragmenta- tion or surface catalyzed secondary nucleation, (3c) cap- tures the buildup of aggregate mass in each compartment due to elongation of existing aggregation by monomer addition, and (3d) models the population balance of monomers in each compartment as a result of aggregate growth and the flux monomer between compartments I and II. This flux is given by ((3a)) ensuring that parti- tioning is maintained close to the monomer enrichment factor Γ. While the monomer enrichment factor Γ ((14)) governs the constant ratio of the time dependent concentrations in compartment I and II, the partitioning degree ξ ((2)) determines how the total monomer concentration, which decays over time as a result of aggregation, is split be- tween the two compartments at any time point during the kinetics of aggregation. As we will see, both parame- ters will be crucial in controlling the degree of aggregate enrichment inside the compartments. IRREVERSIBLE AGGREGATION IN THE PRESENCE OF PHASE SEPARATED COMPARTMENTS To understand how protein aggregation kinetics cou- ples to two phase separated compartments in terms of the physical parameters Γ and ξ, we constructed explicit analytical solutions to the set of non-linear kinetic equa- tions (3) by exploiting an analogy to classical mechan- ics ([46] and SI, section S3, for details), and compared with numerical solutions of (3). Monomer enrichment causes a relative change in nucleation and growth of aggregates between the compartments In the limit of fast monomer diffusion the aggregation kinetics in each compartment is controlled by a set of ef- fective rate parameters. The relative magnitude of these effective rates between compartment I and II at early times scales with the monomer enrichment as Γn1 , while at late times, the corresponding ratio of these rates scales with Γ(n2+1)/2 (see SI, Eq. (S33) and (S34)). Thus, the aggregate growth inside compartment I is faster than in compartment II if there is enrichment of monomers in the condensed phase, i.e., when Γ > 1. Moreover, the relative magnitudes of growth rate at early times solely depends on the reaction order of primary nucleation, n1, while at late times, relative growth is determined by the reaction order of secondary nucleation, n2. Phase separated compartments mediate a positive feedback for aggregate growth This difference in growth rates between the phases can be qualitatively explained by the rapid preference of monomers to recover phase equilibrium (Fig. 3(a)). The higher monomer concentration in compartment I causes aggregates to nucleate first inside compartment I. As a consequence elongation of aggregates is more pronounced inside compartment I leading to a stronger consump- tion of monomers. This difference in monomer consump- 4 tion between the compartments couples to the flux (3), which forces more monomers to diffuse into compartment I to maintain partitioning equilibrium, even as aggregates grow. This positive feedback mechanism in compartment I is accompanied by negative feedback for compartment II, which continuously looses monomers leading to a slow down of the aggregation kinetics outside. Thus, the cou- pling between the aggregation kinetics and phase sepa- ration, mediated by (3), is key to determine aggregate enrichment/depletion in each phase. Positive feedback for aggregate growth causes strong aggregate enrichment a(t)/cII a(t) and cII To understand this feedback mechanism we study the time evolution of the aggregate concentration inside each phase, cI a (t) (Fig. 3(b)). As aggregation is ini- tiated by primary nucleation, it is solely determined by the monomer concentration. Because monomer concen- trations in the compartments are slaved due to the rapid flux that maintains partitioning equilibrium, the time evolution of the aggregate concentrations in the early regime of the aggregation kinetics are slaved as well, fol- a (t) ∝ Γn1. When aggregates start con- lowing cI suming monomers via elongation, the flux of monomers from compartment II to I causes a saturation of the ag- gregate concentration outside the compartment II, while the concentration of aggregates in compartment I in- creases significantly. This rapid increase of growth is facilitated by the continuous influx of monomers (pos- itive feedback). As monomers get depleted in the entire system the growth of aggregates also saturates in com- partment I. Most importantly, the resulting asymptotic concentrations at large time scales, cI a (∞), can differ by several orders of magnitude, even for modest values of Γ corresponding to weak relative interactions. a(∞) and cII Enrichment and depletion relative to homogeneous aggregation is determined by the reaction orders To elucidate the impact of the reactions orders on the aggregation kinetics we first consider the enrichment of aggregates relative to the case of homogeneous aggrega- tion, i.e., for Γ = 1. For large values of monomer enrich- ment, the asymptotic concentrations in compartments I and II at large times relative to the homogeneous aggre- gate concentration c0 a at large times (see SI, Eq. (S46), for the definition) read 2 Γ cII aw (cid:39) ξ( ¯φ)n1− n2+1 a (∞) (cid:0)ξ( ¯φ) Γ(cid:1) n2+1 c0 cI a(∞) a (cid:39) c0 2 − n2+1 2 , , (4) (5) where w is a dimensionless numerical prefactor (see SI, Eq. (S49)). We see that for a large monomer enrich- 5 FIG. 3. Segregation of aggregates into compartment I via positive feedback mediated by phase separation. (a) Sketch of aggregation kinetics inside the two compartments I and II. Left: Initially, monomers get enriched on a short diffusive time scales due to the partitioning mediated by the phase separated compartments ((14)). Center: Monomers slowly aggregate. More aggregates nucleate and grow in compartment I due to the initial partitioning of monomers. This pronounced, initial aggregation causes a continuous monomer flux into compartment I, further promoting aggregation (positive feedback indicated by arrows). Right: Partitioning of monomers together with the positive feedback can cause a very pronounced accumulation of aggregates relative to compartment II. (b) Aggregate concentration cα a (t) as a function of time t obtained from solving numerically and analytically Eqs. (3) actually confirms that aggregates can enrich by several orders of magnitude. (c) The a (∞) inside each of the compartment inversely scale for small compartments, while for asymptotic concentrations cI large compartment I, aggregate enrichment therein vanishes while depletion inside compartment II is dominated by primary nucleation.The asymptotic concentration in the absence of monomer enrichment, Γ = 1, is denoted as c0 a. Dashed line are the scalings given in the the main text. Parameters: n1 = n2 = 2. (d) Enrichment factor ε of aggregates inside compartment I as a function of monomer enrichment Γ can reach very large values. The behaviour switches from secondary nucleation dominated increase at small compartment I volumes to primary dominated growth at large volumes. Dashed line are the scalings given in (6). (e) The slope of the enrichment factor as a function of mean volume fraction ¯φ, equivalently speaking, volume of compartment I, changes its sign when enrichment is dominated by primary (n1 = 2, n2 = 0) or secondary nucleation (n1 = 2, n2 = 2). Parameters: (b,e) Γ = 3 consistent with weak interactions. a(∞) and cII ment factor Γ, the enrichment of aggregates inside com- partment I gets more pronounced, while aggregates in compartment II are more depleted relative to the homo- geneous case (Fig. 3(c)). Most impotantly, the value of the aggregate concentration for given monomer enrich- ment is controlled by the reaction orders for primary and secondary nucleation, n1 and n2. Aggregate concentration in the compartments is controlled by compartment volume Having understood the role of the monomer enrich- ment factor Γ in aggregation kinetics, we now turn to how the asymptotic concentrations of aggregates in each compartment depend on the volumes of the compart- ments. The dependence on compartment volume is given the partition degree ξ( ¯φ). From (2), we see that the par- tition degree ξ( ¯φ) ∈ [1, Γ−1], where the value of one is relevant for small compartments (Fig. 2(b)). Following a(∞) ∝ a (∞)(cid:1)−1 (cid:0)cII (4) and (5), we see that for a small volume of compart- ment I, enrichment and depletion exhibit an inverse scal- ing with cI , which is solely dependent on the reaction order for secondary nucleation. Contrariwise, when the volume of compartment I is large, enrichment of aggregates inside I vanishes, while deple- tion inside compartment II then solely depends on the reaction order for primary nucleation, cII a (∞) ∝ Γ−n1. n2+1 2 ∝ Γ This switch between aggregate enrichment governed by secondary nucleation, to an enrichment solely deter- mined by primary nucleation, arises from primary nucle- ation events occurring first inside compartment I due to a higher monomer concentration (Γ > 1). Once the first aggregates have formed via primary nucleation inside compartment I, small and large compartments behave fundamentally differently. If compartment I is small, only a few aggregates can form via primary nucleation due to the small compartment size. As aggregates be- gin to grow earlier in compartment I, the unbalance of monomers causes a flux from II to I. As a consequence aggregateenrichment"monomerenrichment100102104106108100101102103(d)c↵a/cIa(1)timet0104102100102101100101(b)IIIaggregateenrichment"/"(II)volumefraction¯01200.51(e)primarysecondaryIIIIIIIIIIIItimetaggregateconc.c↵a/cIa(1)timet/⌧104102100102104106(b)III(a)monomerpartitioningIIIphaseseparationreplenishesconcentrationinsideaggregatepartitioningc↵a(1)/c0amonomerenrichment106104102100102104100101102103(c)IIImonomerenrichment<latexit sha1_base64="cVdJa1KnaUeiB5kyvAxFM0Zp8KM=">AAACCnicbVC7SgNBFJ2NrxhfUUub0SBYhV0RtAxaaBnBPCBZwuzkJhkyj2VmVgxLaht/xcZCEVu/wM6/cZJsoYkHBg7n3Hvn3hPFnBnr+99ebml5ZXUtv17Y2Nza3inu7tWNSjSFGlVc6WZEDHAmoWaZ5dCMNRARcWhEw6uJ37gHbZiSd3YUQyhIX7Ieo8Q6qVM8bFt4sKlQUgnQGKRmdCBAWjzG7WsiBOkUS37ZnwIvkiAjJZSh2il+tbuKJpMhlBNjWoEf2zAl2jLKYVxoJwZiQoekDy1HJRFgwnR6yhgfO6WLe0q755aYqr87UiKMGYnIVQpiB2bem4j/ea3E9i7ClMk4sSDp7KNewrFVeJIL7jIN1PKRI4Rq5nbFdEA0odalV3AhBPMnL5L6aTnwy8HtWalymcWRRwfoCJ2gAJ2jCrpBVVRDFD2iZ/SK3rwn78V79z5mpTkv69lHf+B9/gCpEJrX</latexit><latexit sha1_base64="cVdJa1KnaUeiB5kyvAxFM0Zp8KM=">AAACCnicbVC7SgNBFJ2NrxhfUUub0SBYhV0RtAxaaBnBPCBZwuzkJhkyj2VmVgxLaht/xcZCEVu/wM6/cZJsoYkHBg7n3Hvn3hPFnBnr+99ebml5ZXUtv17Y2Nza3inu7tWNSjSFGlVc6WZEDHAmoWaZ5dCMNRARcWhEw6uJ37gHbZiSd3YUQyhIX7Ieo8Q6qVM8bFt4sKlQUgnQGKRmdCBAWjzG7WsiBOkUS37ZnwIvkiAjJZSh2il+tbuKJpMhlBNjWoEf2zAl2jLKYVxoJwZiQoekDy1HJRFgwnR6yhgfO6WLe0q755aYqr87UiKMGYnIVQpiB2bem4j/ea3E9i7ClMk4sSDp7KNewrFVeJIL7jIN1PKRI4Rq5nbFdEA0odalV3AhBPMnL5L6aTnwy8HtWalymcWRRwfoCJ2gAJ2jCrpBVVRDFD2iZ/SK3rwn78V79z5mpTkv69lHf+B9/gCpEJrX</latexit><latexit sha1_base64="cVdJa1KnaUeiB5kyvAxFM0Zp8KM=">AAACCnicbVC7SgNBFJ2NrxhfUUub0SBYhV0RtAxaaBnBPCBZwuzkJhkyj2VmVgxLaht/xcZCEVu/wM6/cZJsoYkHBg7n3Hvn3hPFnBnr+99ebml5ZXUtv17Y2Nza3inu7tWNSjSFGlVc6WZEDHAmoWaZ5dCMNRARcWhEw6uJ37gHbZiSd3YUQyhIX7Ieo8Q6qVM8bFt4sKlQUgnQGKRmdCBAWjzG7WsiBOkUS37ZnwIvkiAjJZSh2il+tbuKJpMhlBNjWoEf2zAl2jLKYVxoJwZiQoekDy1HJRFgwnR6yhgfO6WLe0q755aYqr87UiKMGYnIVQpiB2bem4j/ea3E9i7ClMk4sSDp7KNewrFVeJIL7jIN1PKRI4Rq5nbFdEA0odalV3AhBPMnL5L6aTnwy8HtWalymcWRRwfoCJ2gAJ2jCrpBVVRDFD2iZ/SK3rwn78V79z5mpTkv69lHf+B9/gCpEJrX</latexit><latexit sha1_base64="cVdJa1KnaUeiB5kyvAxFM0Zp8KM=">AAACCnicbVC7SgNBFJ2NrxhfUUub0SBYhV0RtAxaaBnBPCBZwuzkJhkyj2VmVgxLaht/xcZCEVu/wM6/cZJsoYkHBg7n3Hvn3hPFnBnr+99ebml5ZXUtv17Y2Nza3inu7tWNSjSFGlVc6WZEDHAmoWaZ5dCMNRARcWhEw6uJ37gHbZiSd3YUQyhIX7Ieo8Q6qVM8bFt4sKlQUgnQGKRmdCBAWjzG7WsiBOkUS37ZnwIvkiAjJZSh2il+tbuKJpMhlBNjWoEf2zAl2jLKYVxoJwZiQoekDy1HJRFgwnR6yhgfO6WLe0q755aYqr87UiKMGYnIVQpiB2bem4j/ea3E9i7ClMk4sSDp7KNewrFVeJIL7jIN1PKRI4Rq5nbFdEA0odalV3AhBPMnL5L6aTnwy8HtWalymcWRRwfoCJ2gAJ2jCrpBVVRDFD2iZ/SK3rwn78V79z5mpTkv69lHf+B9/gCpEJrX</latexit>"/"(II)aggregateenrichment<latexit sha1_base64="jy2YgOgUN4Dqcu/1GckM7Ya3npI=">AAACBnicbVC7SgNBFJ2Nrxhfq5YiDAbBKuyKoGXQxjKCSYRkCbOTm82Q2Qczd8WwpLLxV2wsFLH1G+z8G2eTLTTxwMDhnHu4c4+fSKHRcb6t0tLyyupaeb2ysbm1vWPv7rV0nCoOTR7LWN35TIMUETRRoIS7RAELfQltf3SV++17UFrE0S2OE/BCFkRiIDhDI/Xswy7CA2YsCBQEDIFCpAQfhhAhndCeXXVqzhR0kbgFqZICjZ791e3HPM3jXDKtO66ToJcxhYJLmFS6qYaE8RELoGNoxELQXjY9Y0KPjdKng1iZZ9ZP1d+JjIVaj0PfTIYMh3rey8X/vE6KgwsvE1GSIkR8tmiQSooxzTuhfaGAoxwbwrgS5q+UD5liHE1zFVOCO3/yImmd1lyn5t6cVeuXRR1lckCOyAlxyTmpk2vSIE3CySN5Jq/kzXqyXqx362M2WrKKzD75A+vzB3z2mRo=</latexit><latexit sha1_base64="jy2YgOgUN4Dqcu/1GckM7Ya3npI=">AAACBnicbVC7SgNBFJ2Nrxhfq5YiDAbBKuyKoGXQxjKCSYRkCbOTm82Q2Qczd8WwpLLxV2wsFLH1G+z8G2eTLTTxwMDhnHu4c4+fSKHRcb6t0tLyyupaeb2ysbm1vWPv7rV0nCoOTR7LWN35TIMUETRRoIS7RAELfQltf3SV++17UFrE0S2OE/BCFkRiIDhDI/Xswy7CA2YsCBQEDIFCpAQfhhAhndCeXXVqzhR0kbgFqZICjZ791e3HPM3jXDKtO66ToJcxhYJLmFS6qYaE8RELoGNoxELQXjY9Y0KPjdKng1iZZ9ZP1d+JjIVaj0PfTIYMh3rey8X/vE6KgwsvE1GSIkR8tmiQSooxzTuhfaGAoxwbwrgS5q+UD5liHE1zFVOCO3/yImmd1lyn5t6cVeuXRR1lckCOyAlxyTmpk2vSIE3CySN5Jq/kzXqyXqx362M2WrKKzD75A+vzB3z2mRo=</latexit><latexit sha1_base64="jy2YgOgUN4Dqcu/1GckM7Ya3npI=">AAACBnicbVC7SgNBFJ2Nrxhfq5YiDAbBKuyKoGXQxjKCSYRkCbOTm82Q2Qczd8WwpLLxV2wsFLH1G+z8G2eTLTTxwMDhnHu4c4+fSKHRcb6t0tLyyupaeb2ysbm1vWPv7rV0nCoOTR7LWN35TIMUETRRoIS7RAELfQltf3SV++17UFrE0S2OE/BCFkRiIDhDI/Xswy7CA2YsCBQEDIFCpAQfhhAhndCeXXVqzhR0kbgFqZICjZ791e3HPM3jXDKtO66ToJcxhYJLmFS6qYaE8RELoGNoxELQXjY9Y0KPjdKng1iZZ9ZP1d+JjIVaj0PfTIYMh3rey8X/vE6KgwsvE1GSIkR8tmiQSooxzTuhfaGAoxwbwrgS5q+UD5liHE1zFVOCO3/yImmd1lyn5t6cVeuXRR1lckCOyAlxyTmpk2vSIE3CySN5Jq/kzXqyXqx362M2WrKKzD75A+vzB3z2mRo=</latexit><latexit sha1_base64="jy2YgOgUN4Dqcu/1GckM7Ya3npI=">AAACBnicbVC7SgNBFJ2Nrxhfq5YiDAbBKuyKoGXQxjKCSYRkCbOTm82Q2Qczd8WwpLLxV2wsFLH1G+z8G2eTLTTxwMDhnHu4c4+fSKHRcb6t0tLyyupaeb2ysbm1vWPv7rV0nCoOTR7LWN35TIMUETRRoIS7RAELfQltf3SV++17UFrE0S2OE/BCFkRiIDhDI/Xswy7CA2YsCBQEDIFCpAQfhhAhndCeXXVqzhR0kbgFqZICjZ791e3HPM3jXDKtO66ToJcxhYJLmFS6qYaE8RELoGNoxELQXjY9Y0KPjdKng1iZZ9ZP1d+JjIVaj0PfTIYMh3rey8X/vE6KgwsvE1GSIkR8tmiQSooxzTuhfaGAoxwbwrgS5q+UD5liHE1zFVOCO3/yImmd1lyn5t6cVeuXRR1lckCOyAlxyTmpk2vSIE3CySN5Jq/kzXqyXqx362M2WrKKzD75A+vzB3z2mRo=</latexit>volumefraction¯<latexit sha1_base64="FwvLbQ591lc7iWe1fYnVQ6GVFHg=">AAACC3icbVDLSsNAFJ3UV62vqEs3Q4vgqiQi6LLoxmUF+4AmlMl00g6dTMLMTbGE7t34K25cKOLWH3Dn3zhps9DWAxcO59w7c+8JEsE1OM63VVpb39jcKm9Xdnb39g/sw6O2jlNFWYvGIlbdgGgmuGQt4CBYN1GMRIFgnWB8k/udCVOax/IepgnzIzKUPOSUgJH6dtUD9gDZJBZpxHCoCM11PMNeQBT2khHHfbvm1J058CpxC1JDBZp9+8sbxNS8J4EKonXPdRLwM6KAU8FmFS/VLCF0TIasZ6gkEdN+Nr9lhk+NMsBhrExJwHP190RGIq2nUWA6IwIjvezl4n9eL4Xwys+4TFJgki4+ClOBIcZ5MHjAFaMgpoYQqrjZFdMRyQMx8VVMCO7yyaukfV53nbp7d1FrXBdxlNEJqqIz5KJL1EC3qIlaiKJH9Ixe0Zv1ZL1Y79bHorVkFTPH6A+szx/E95rU</latexit><latexit sha1_base64="FwvLbQ591lc7iWe1fYnVQ6GVFHg=">AAACC3icbVDLSsNAFJ3UV62vqEs3Q4vgqiQi6LLoxmUF+4AmlMl00g6dTMLMTbGE7t34K25cKOLWH3Dn3zhps9DWAxcO59w7c+8JEsE1OM63VVpb39jcKm9Xdnb39g/sw6O2jlNFWYvGIlbdgGgmuGQt4CBYN1GMRIFgnWB8k/udCVOax/IepgnzIzKUPOSUgJH6dtUD9gDZJBZpxHCoCM11PMNeQBT2khHHfbvm1J058CpxC1JDBZp9+8sbxNS8J4EKonXPdRLwM6KAU8FmFS/VLCF0TIasZ6gkEdN+Nr9lhk+NMsBhrExJwHP190RGIq2nUWA6IwIjvezl4n9eL4Xwys+4TFJgki4+ClOBIcZ5MHjAFaMgpoYQqrjZFdMRyQMx8VVMCO7yyaukfV53nbp7d1FrXBdxlNEJqqIz5KJL1EC3qIlaiKJH9Ixe0Zv1ZL1Y79bHorVkFTPH6A+szx/E95rU</latexit><latexit sha1_base64="FwvLbQ591lc7iWe1fYnVQ6GVFHg=">AAACC3icbVDLSsNAFJ3UV62vqEs3Q4vgqiQi6LLoxmUF+4AmlMl00g6dTMLMTbGE7t34K25cKOLWH3Dn3zhps9DWAxcO59w7c+8JEsE1OM63VVpb39jcKm9Xdnb39g/sw6O2jlNFWYvGIlbdgGgmuGQt4CBYN1GMRIFgnWB8k/udCVOax/IepgnzIzKUPOSUgJH6dtUD9gDZJBZpxHCoCM11PMNeQBT2khHHfbvm1J058CpxC1JDBZp9+8sbxNS8J4EKonXPdRLwM6KAU8FmFS/VLCF0TIasZ6gkEdN+Nr9lhk+NMsBhrExJwHP190RGIq2nUWA6IwIjvezl4n9eL4Xwys+4TFJgki4+ClOBIcZ5MHjAFaMgpoYQqrjZFdMRyQMx8VVMCO7yyaukfV53nbp7d1FrXBdxlNEJqqIz5KJL1EC3qIlaiKJH9Ixe0Zv1ZL1Y79bHorVkFTPH6A+szx/E95rU</latexit><latexit sha1_base64="FwvLbQ591lc7iWe1fYnVQ6GVFHg=">AAACC3icbVDLSsNAFJ3UV62vqEs3Q4vgqiQi6LLoxmUF+4AmlMl00g6dTMLMTbGE7t34K25cKOLWH3Dn3zhps9DWAxcO59w7c+8JEsE1OM63VVpb39jcKm9Xdnb39g/sw6O2jlNFWYvGIlbdgGgmuGQt4CBYN1GMRIFgnWB8k/udCVOax/IepgnzIzKUPOSUgJH6dtUD9gDZJBZpxHCoCM11PMNeQBT2khHHfbvm1J058CpxC1JDBZp9+8sbxNS8J4EKonXPdRLwM6KAU8FmFS/VLCF0TIasZ6gkEdN+Nr9lhk+NMsBhrExJwHP190RGIq2nUWA6IwIjvezl4n9eL4Xwys+4TFJgki4+ClOBIcZ5MHjAFaMgpoYQqrjZFdMRyQMx8VVMCO7yyaukfV53nbp7d1FrXBdxlNEJqqIz5KJL1EC3qIlaiKJH9Ixe0Zv1ZL1Y79bHorVkFTPH6A+szx/E95rU</latexit> totic aggregate enrichment ratio ε( ¯φ) = cI a (∞) ∝ ξ( ¯φ)n2−n1+1 Γn2+1 . a(∞) cII 6 (6) As the compartment volume enters the enrichment fac- tor ε( ¯φ) solely via the partition degree ξ( ¯φ), the sign of n2 − n1 + 1 determines whether larger or smaller com- partments lead to a larger enrichment (Fig. 3(e)). Indeed we find that the slope of the enrichment factor scales as ε( ¯φ)(cid:48) ∝ (n1 − n2 − 1). Thus, for n1 > n2 + 1, increas- ing the compartment volume by increasing the amount of A-material ¯φ causes a larger relative enrichment. Con- versely, for n1 < n2 + 1, larger enrichment can be found for smaller compartment sizes. Consistently, if the nucle- ation coefficients obey n2 = n1 − 1, compartment volume has no impact on the enrichment factor ε. This qualitative switch in the mechanism for aggregate enrichment raises the question which systems favor large or small compartment volumes in order to maximize ag- gregate enrichment ε( ¯φ). Figure 4 depicts the regimes in terms of the reaction orders characterizing primary and secondary nucleation, n1 and n2, for which the maximal aggregate enrichment corresponds to smaller and larger compartment volumes. This prediction can be related to specific aggregating systems for which the values of the reaction orders n1 and n2 have been experimentally de- termined (References see caption of Fig. 4). Using these values for the reaction orders, our model predicts that largest enrichment is obtained for large compartments in systems of aggregating tau and yeast prion Ure2p. These two examples belong primarily to the class of systems where the mechanism responsible for the formation of new aggregates in the late stage is fragmentation which has a zero secondary reaction order, n2 = 0 (i.e., nu- cleation is monomer independent). For non-fragmenting systems with n2 > 0, our model predicts different sce- narios for aggregating systems: largest aggregate en- richment for large compartment volumes occurs in the case branching systems, such as actin in the presence of the complex Arp2/3, as well as systems proliferat- ing through monomer dependent secondary nucleation with n2 < n1 − 1, such as the Islet Amyloid Polypep- tide (IAPP). In contrast, largest aggregate enrichment is reached for small compartments in the case of the 40- and 42-residue forms of Amyloid-β peptide (Aβ40 and Aβ42). CONCLUSION By combining the theories of irreversible protein ag- gregation kinetics and phase separation we have shown how liquid compartments can control the spatial pattern of irreversible protein aggregation. The coupling of slow aggregation and rapid phase separation leads to a mech- anism whereby even a weak partitioning of monomers in the case of weak protein-protein interactions is am- plified into a relatively large accumulation of aggregates FIG. 4. Theoretical predictions of maximal aggregate enrichment for various aggregating systems. Our pre- dictions are summarized by a phase diagram depicting that aggregating systems characterized by different reaction orders for primary and secondary nucleation, n1 and n2, show maxi- mal aggregate enrichment for large or small compartments, respectively. The two regions where either large or small compartments lead to a larger enrichment of aggregates is separated by the line n2 = n1 − 1 determined from (6). For n2 > n1 − 1 small compartments lead to larger aggregate en- richment, while for n2 < n1−1, larger compartments are ben- eficial. To illustrate which scenario might apply to which kind of aggregating system, we indicate the measured values of the primary and secondary reaction orders for a range of systems propagating through fragmentation (blue), lateral branching (green) or monomer-dependent secondary nucleation (red): Tau [48], yeast prion Ure2p [49], IAPP [50], amyloid-β40 (for monomer concentrations below 5 µM) [51], amyloid-β42 [41]. of this continuous flux, the secondary nucleation events quickly overwhelm primary nucleation events inside com- partment I, while secondary nucleation is suppressed in compartment II. However, if compartment I is large, the aggregation kinetics is similar to that for a homogeneous system because the monomer mass concentration is very close to the total monomer mass in the system and there is only a negligible amount of monomers entering from compartment II. Additionally, in the smaller compart- ment II where aggregates grow via primary nucleation, the coupling flux continuously removes monomers sup- pressing primary nucleation. Since compartment I is large, it shows little or no enrichment of aggregates rela- tive to the homogeneous case while inside the small com- partment II, aggregates are depleted determined by the lack of primary nucleation events relative to the homo- geneous case. Changes in compartment volume switch the driving mechanism for aggregate enrichment To quantify the switch in aggregation enrichment as a function of compartment volume we define the asymp- Aβ40, Aβ42Yeast prionIAPPTau0123456780123456Primarynucleationreactionordern1Secondarynucleationreactionordern2Actin + Arp2/3FragmentationBranchingMonomer-dependent secondary nucleation 7 In summary, our model suggests that liquid compart- ments have a propensity to enrich aggregates and thus could play a role in controlling them spatially. In our model we have considered the case when monomers and aggregates do not affect phase separation and when phase separation is driven by the competition between the entropic tendency to mix and interactions favoring demixing. Future work could be devoted to extending our model by entropically driven phase separation, relevant for the assembly of coacervates [59] or mixtures with depletion interactions, or to including in our model a coupling between aggregates and the liquid compart- ment. Moreover, our model is restricted to time scales when aggregates hardly diffuse. Diffusion of aggregates may diminish the observed strong aggregate enrichment suggesting new avenues for further studies. A lowered enrichment could also be caused by the coarsening dynamics of many droplets [60 -- 62]. While coarsening via coalescence would not affect our results at all because aggregates remain confined inside the droplets, we leave it to future work to understand to which extent dissolving droplets undergoing Ostwald ripening would diminish the degree of aggregate enrichment. Acknowledgements CAW thanks the German Research Foundation (DFG) for financial support and TCTM acknowledges the sup- port from the Swiss National Science foundation. in the compartment. The resulting degree of aggregate enrichment is determined by two physical parameters re- lated to phase separation (monomer enrichment Γ and partitioning degree ξ) and two parameters characteriz- ing the aggregation kinetics (reaction orders n1 and n2 for primary and secondary nucleation). Since the kinetic parameters are fixed by the nature of the nucleation reac- tions, the phase separation parameters that are governed by the molecular interactions between the constituents thus represent ideal control variables to regulate the de- gree of aggregate enrichment. Our model may already provide a framework to explain the phenomena of aggregate partitioning inside living cells. An example of such phenomena could be the par- titioning of pericentriolar material into centrosomes [52] and the spatial organization of aggregates inside stress granules [19, 22]. The propensity of aggregates to so- lidify the compartment as reported in Ref. [22] could be accounted for in our model through a gel-sol tran- sition [53, 54]. Including the solidification induced by aggregates could lead to additional volume changes of the compartment which in turn may affect the aggregation kinetics. Our quantitative predictions of strong aggregate en- richment inside a liquid compartment (Fig. 3 (b-e)) and how compartment volume affects this enrichment for dif- ferent aggregating systems (Fig. 4) suggest direct experi- mental tests for instance using an Amyloid-β aggregation assay [55] in systems where one can enhance the thermo- dynamic stability of droplets using surfactant-stabilized emulsions. The condition of negligible aggregate diffu- sion can be realized using a gel matrix with pore sizes well above the monomer size but much lower than the expected average fibril length. The monomer enrichment Γ could be varied in vitro by changing parameters such as salt composition, pH or temperature, however, ver- ifying that these changes only weakly affect the phase separation itself. Furthermore, the enrichment of toxic aggregates inside liquid compartments suggests new directions for drug de- sign against aberrant protein aggregation. Our results suggest to design drugs not only with respect to their ability to interfere with the aggregation kinetics [56] but also with respect to their partitioning properties into the liquid compartments. This strategy is reminiscent of quantifying the potency of low molecular weighted anaes- thetics by the Meyer-Overton correlation [57, 58]. The reported feedback mechanism of aggregate growth mediated by liquid compartments may represent a gen- eral principle to spatially confine and speed up other irre- versible chemical processes. Examples may include pre- cipitation of proteins or polymerization kinetics of actin and microtubules (see also Fig. 4). Indeed, a speed up of the chemical reactions could be expected due to the in- creased concentration of educts inside the liquid compart- ments. Thus liquid compartments are ideal biomolecular microreactors that enrich the amount of products by dy- namically exchanging reactants with their surroundings. SUPPLEMENTAL INFORMATION I. PARTITION COEFFICIENT FOR DILUTE MONOMERS AT EQUILIBRIUM At equilibrium the chemical potentials of each component inside (I) and outside (II) the compartment are balanced leading to relationships between the concentration values inside and outside. Specifically, if phase separation equi- librium is reached the following conditions are fulfilled 8 We consider phase separation of an incompressible, ternary mixture composed of monomers, component A and B (we neglect the interactions between the aggre- gates and phase separation for simplicity) described by the following Flory-Huggins free energy density [63, 64] ln(φA) + φB νB ln(φB) + cm ln(νmcm) (cid:21) + Λ φAφB + cmνm (ΛmAφA + ΛmBφB) , (7) (cid:20) φA νA f = kBT where the logarithmic terms correspond to the mixing entropy. The interactions between A and B are de- scribed by the parameter Λ while the interactions with the monomers are characterized by Λmi, i = A, B. These interaction parameters have the unit [1/volume]. Here we define a dimensionless interaction parameters by writing Λ = χ/ν and Λmi = χmi/ν with ν = νA. Moreover, for simplicity, we consider equal molecular volumes of A and B, ν = νB. Thus we arrive at φA ln(φA) + φB ln(φB) + νcm ln(νmcm) (cid:20) f = kBT ν (cid:21) + χ φAφB + cmνm (χmAφA + χmBφB) . (8) With φB = 1 − φA − νmcm and monomers being dilute, νmcm (cid:28) 1, we can expand f in νmcm up to the first order: φ ln(φ) + (1 − φ) ln(1 − φ) + χ φ(1 − φ) f (φ, cm) (cid:39) kBT ν (cid:26) (cid:18) + νmcm (ν/νm) ln(νmcm) − ln(1 − φ) (cid:19)(cid:27) + (χmA − χmB − χ) φ + χmB − 1 , (9) where we defined φA ≡ φ and neglected all term of or- der O . The corresponding chemical potentials read (cid:18) ln φ − ln (1 − φ) + χ(1 − 2φ) + cmνm χmA − χ − χmB + 1 1 − φ (10a) (cid:19)(cid:21) , νm ν ln(1 − φ) (10b) (cid:21)(cid:27) ln (νmcm) + 1 + + νm ν (χmA − χ) φ + χmB(1 − φ) − 1 . µm(φ, cm) = ∂f ∂cm = kBT (cid:26) (cid:20) (cid:104) (νmcm)2(cid:105) µ(φ, cm) = ν ∂f ∂φ (cid:20) = kBT µI(φI, cI µI m(φI, cI m) = µII(φII, cII m) = µII m(φII, cII m) , m) . (11a) (11b) m and cII The relations above allow to calculate the equilibrium concentration in each phase, for component A, φI and φII, and the monomers, cI m. An analytic result of the equilibrium concentrations is very difficult to obtain. However, we can focus on the leading contributions for the balance of the chemical potentials inside and outside taking advantage that monomers are dilute and thereby obtain an approximation for the equilibrium values inside and outside. Considering that the dimensionless interaction param- eters χ are all of O(1), the impact of the dilute monomers on the phase equilibrium between A and B is negligible and we can approximate (cid:20) (cid:18) φ (cid:19) 1 − φ (cid:21) µ(φ, cm) (cid:39) kBT ln + χ(1 − 2φ) . (12) The resulting chemical potential Eq. (12) simply corre- sponds to the chemical potential of a binary, incompress- ible Flory-Huggins mixture. From the equilibrium condi- tion Eq. (11a), we can calculate the binodal line described by the condition χ (cid:39) ln (φ/(1 − φ)) 2φ − 1 , (13) which solely depends on the interaction parameters be- tween A and B, χ. By means of equilibrium condition Eq. (11b), we can calculate the impact of the phase sepa- rated compartment on the monomer distribution leading to the monomer enrichment (cid:104) νm ν ∆χ(cid:0)φI − φII(cid:1)(cid:105) cI m m (cid:39) exp cII =: Γ , (14) where the relative interaction strength reads ∆χ = (χB,m − χA,m). Thus, there is enrichment of monomers in the condensed phase (Γ > 1) if monomers favor the presence of A relative to B, i.e., χA,m < χB,m. Inter-compartment flux of monomers close to equilibrium This flux can be calculated for a maintained concen- tration difference φI−φII using the chemical potential for the monomers µm (Eq. (10b)). To this end, let us assume that the compartment is spherical with a radius R and corresponding volume V I = (4π/3)R3. Perturbing the II. ANALYTICAL SOLUTION TO AGGREGATION KINETICS WITH LIQUID COMPARTMENTS 9 (15) In this appendix, we discuss the detail associated with the derivation of analytical solutions to the aggregation kinetics in the presence of liquid compartments, Eq.(3), main text. concentrations in both phases may lead to an unbalance of the chemical potential and thus a flux between the phases. The total change of monomer due to a net flux through the interface between the compartments reads: (cid:90) dϕ dθ er · j(cid:12)(cid:12)R−∆x/2 R+∆x/2 , J = R2 lim ∆x→0 where the unbalance of the chemical potential occurs through the interface of width ∆x and er is the radial unit vector pointing normal to the spherical interface. The local flux j can be approximated as j = −ξ∇µm (cid:39) −ξer∂rµm (cid:39) −erξ(µI m)/∆x, where ξ denotes the mobility coefficient. This approximation neglects spatial inhomogeneities in chemical potentials within the phases which is valid in the limit of fast diffusion on the length scale of the system of volume V . Using the chemical po- tential of the monomers, Eq. (10b), we can approximate the gradient of the chemical potential as m − µII m − µII µI ∆x m kBT ∆x kBT ∆x = (cid:39) (cid:2)ln(cid:0)M I m (cid:1) − ln(cid:0)Γ M II m (cid:1)(cid:3) δM I m − Γ δM II M I m m,eq , (16) (17) m = M I m,eq + δM I m,eq + δM II where we expanded M I m = M II m up to linear order around the equilib- rium concentrations M I m,eq/Γ, re- spectively. Thus the change in concentration due to the exchange of material through the interface reads m,eq and M II m and M II m,eq = M I J V (α)(R) (cid:39) − M I 4πR2 Dm ∆x V (α)(R) = −k(α)(R)(cid:0)M I m − ΓM II νm M I m − ΓM II (cid:1) , m,eq m m (18) (19) where the diffusion constant Dm = kbT νmξ. To ease notation we omitted the "δ" to indicate linear deviations from equilibrium. Moreover, the rate to relax back to monomer partitioning at equilibrium is kα(R) = 4πR2Dm ∆xV (α)(R)νmM I m,eq . (20) In particular, This rate depends on parameters such as the monomer diffusion constant and the size of the compartment V I (Eq. (20)). for large compartment I, V I ≈ V , the fraction of rates, kI/kII, decreases toward zero indicating that the relaxation in compartment I is slowed down relative to compartment II. Conversely, in the case of small compartments, V I (cid:28) V , relaxation of compartment I is fast compared to compartment II. Most importantly, the size of the compartment does not affect the equilibrium concentration. For simplicity, we neglect the impact of surface tension which leads to a weak in- crease of the equilibrium volume fractions. A. Boundary layer dynamics Due to the separation of timescales between monomer equilibration between the two compartments and protein aggregation, the system described by Eqs. (3) (main text) will develop initially through a rapid phase of equilibra- tion (boundary layer), before any aggregation occurs in either compartment. During this phase, the initial val- ues of the monomer concentration in each compartment, M I m (0), relax quickly to equilibrium such that the condition M I m (t) is satisfied before aggre- gation is initiated. This early equilibration kinetics is described by setting the aggregation terms in Eqs. (3) (main text) to zero, yielding the following equations: m(0) and M II m(t) = ΓM II dM I a(t) dt dM II a (t) dt m(t) − ΓM II m(t) − ΓM II = −kI(cid:2)M I = kII(cid:2)M I kI(cid:2)M I kII(cid:2)M I m (t)(cid:3) , m (t)(cid:3) . m (0)(cid:3) m (0)(cid:3) where A = kI + ΓkII, and B =(cid:2)kIIM I m(0) − ΓM II m(0) − ΓM II The solution to Eqs. (21) is: m = ΓB + m = B + M II M I A A e e −At , −At , (22b) m (0)(cid:3)/A. m(0) + kIM II (21a) (21b) (22a) m(t) = ΓM II Note that the kinetics described by Eq. (22) 'pushes' the system towards the slow manifold, which is described by M I m (0), there is no initial phase of 'correction' of the initial con- ditions. At the end of this initial boundary layer phase, the monomer concentrations in the two compartments are given by: m (t). Hence, when M I m(0) = ΓM II M I M II m = ΓB , m = B . (23a) (23b) m(0) = ΓM II Since we are not very much interested in this initial phase of redistribution of the initial conditions, in the following we shall assume for simplicity that the initial monomer concentrations in compartments I and II satisfy the re- lationship M I m (0). This assumption does not affect the generality of our results. In fact, if this condition was not satisfied initially, then, according to Eq. (22), rapid equilibration between the two compart- ments would correct these initial conditions, eventually leading to a 'corrected' set of initial conditions that lie in the slow manifold. 10 II to the total concentration of monomers in the system: M I M II m(t) = ξ ΓMm(t) m (t) = ξ Mm(t) , (27a) (27b) where we abbreviated the partitioning degree ξ = V (Γ − 1)V I + V . (28) Thus, using Eqs. (27), we can re-write Eqs. (24) as: dM I a(t) dt dM II a (t) dt dcI a(t) dt dcII a (t) dt = 2k+(ξ Γ)Mm(t)cI a(t) , = 2k+ξMm(t)cII a (t) , (29a) (29b) = k1(ξ Γ)n1Mm(t)n1 + k2(ξ Γ)n2 Mm(t)n2 M I a(t) , (29c) = k1ξn1 Mm(t)n1 + k2ξn2Mm(t)n2M II a (t) . (29d) 1. Early-time dynamics for aggregate number and mass concentrations in compartments I and II Before discussing the full time course of aggregation, it is useful to consider the early-time kinetics of the system, which emerges when the monomers in the system have not been depleted significantly [5]. This limit is obtained by assuming in (29) that the total monomer concentra- tion is constant in time, i.e., Mm(t) ≈ M tot m [5]. This as- sumption transforms the kinetic equations (29) into the following simpler set of linear differential equations: dM I a(t) dt dM II a (t) dt dcI a(t) dt dcII a (t) dt = µI cI a(t) , = µII cII a (t) , = νI + βI M I a(t) , = νII + βII M II a (t) , (30a) (30b) (30c) (30d) FIG. 5. Comparison between the analytical solutions Eqs. (46), (47), (48) (dashed lines) and the numerical solution to Eqs. (3) in the main text (solid lines). The parameters are: k+ = 106 M−1s−1, k1 = 10−4 M−1s−1, k2 = 104 M−2s−1, M tot m = 1 µM, n1 = n2 = 2, Γ = 3, ξ = 1 and kα/κ0 = 100 for α = I,II. Solving aggregation kinetics in the slow manifold After an initial, rapid phase of monomer redistribu- tion through the two compartments, the system enters a slower phase of dynamics, where, at leading order, the system stays on the slow manifold M I m at all times. For simplicity, let us assume that the initial con- centrations of monomers in the two compartments obey the relationship M I m (0) (otherwise there will be a fast equilibration of the initial conditions such that this relationship is satisfied). To describe the aggregation process in the slow manifold, we write M I m (cid:39) 0 for all times in Eqs. (3) (main text) and find: m − ΓM II m(0) = ΓM II m = ΓM II m(t) cI a(t) = − m (t) cII a (t) = − = −2k+ M I = −2k+ M II = k1 M I m(t)n1 + k2 M I (24a) , (24b) , dM I a(t) dt dM II a (t) dt m(t)n2 M I a(t) , (24c) (24d) dM I m(t) dt dM II m (t) dt dcI a(t) dt dcII a (t) dt = k1 M II m (t)n1 + k2 M II m (t)n2 M II a (t) . It is useful to introduce the total monomer concentration in the system as: Mm(t) = V IM I m(t) + V IIM II m (t) V . (25) where we have introduced the parameters: µI = µ0 ξ Γ , νI = ν0(ξ Γ)n1 , βI = β0(ξ Γ)n2 , (31) Note that this concentration may vary in time as aggre- gates are nucleated and grow, however, the total mass concentration and µII = µ0 ξ , νII = ν0 ξn1 , βII = β0 ξn2 , (32) M tot m = Mm(t) + V IM I a(t) + V IIM II a (t) V (26) with is conserved at all times. Using the condition M I m = ΓM II m , we can write the following relationships linking the concentrations of monomers in compartments I and m )n1 , ν0 = k1(M tot m )n2 , β0 = k2(M tot µ0 = 2k+M tot m . (33a) (33b) (33c) monomermassMIm,MIImtimetκ010−810−710−610−501234(a)IIItotalmassMαm+Mαa[10−6]timetκ000.10.20.30.40.501234(b)III The solution to Eqs. (30) subject to the condition that no aggregates are present initially reads M I M I M II M II a(t) m(0) a (t) m (0) = = λ2 I [cosh(κIt) − 1] κ2 I , (34a) λ2 II[cosh(κIIt) − 1] κ2 II , (34b) for the aggregate mass concentrations in compartments I and II, and cI a(t) = cII a (t) = νI sinh(κIt) κI νII sinh(κIIt) κII (35a) (35b) for the aggregate number concentrations in compart- ments I and II. Here, we have introduced the kinetic coefficients λI = λ0 (ξΓ) n1 2 , λII = λ0 ξ n1 2 , (36) κI = κ0 (ξΓ) n2+1 2 , κII = κ0 ξ n2+1 2 , (37) 11 m(t) = ΓM II 1/κI and 1/κII, characterize the early-time aggregation in the two compartments. Since the growth rate in com- partment I, κI, is much larger than that in compartment II, κII, monomers in compartment I will be consumed by aggregation much faster than those in compartment II. However, the relationship M I m (t) must hold at all times. Thus, to compensate the fast aggregation in compartment I, there will be a flux of monomers from compartment II to compartment I. Eventually, the vast majority of monomers will end up as part of aggregates in compartment I and the parameter κI will naturally con- trol the depletion of monomers in both compartments. We can make this argument more quantitative by using Eqs. (34) as follows. Monomers in compartment I are consumed over a timescale of the order 1/κI. The amount of aggregate mass that will be formed in compartment II during this time period will be of the order M II a (cid:39) M II m (0) λ2 II[cosh(κII/κI) − 1] κ2 II . (41) Since κII/κI (cid:28) 1, we can expand the cosh function as a Taylor series, cosh x = 1 + x2/2 + O(x5). At leading order, we find: M II a (cid:39) M II m (0) λ2 II 2κ2 I . (42) and with (cid:112) (cid:112) λ0 = κ0 = 2k+k1(M tot 2k+k2(M tot m )n1 , m )n2+1 , (38a) (38b) Thus, the ratio between the mass of aggregates formed in compartments I and II over a timescale 1/κI is being the effective rates characterizing the proliferation of aggregates due to primary and secondary nucleation, respectively [5]. According to Eqs. (34) and (35), the aggregate number and mass concentrations in both com- partments grow exponentially with time. The effective growth rates κI and κII are different for each compart- ment and depend on Γ and ξ. Since Γ (cid:29) 1, aggregate growth in the early times is much faster in compartment I compared to compartment II. In particular, the ratio of the growth rates in the two compartments is independent of ξ and is given by: M I a a (cid:39) M II M I M II a(0) a (0) λ2 I II (cid:39) Γ λ2 n1 2 +1. (43) Since Γ (cid:29) 1, the aggregate mass in compartment I will be much larger than that in compartment II. We can thus neglect at leading order the contribution from M II a (t) to the conservation of total mass relationship. Doing so, and using Eqs. (27), we can write the conservation of mass relationship as follows a(t) =(cid:2)M I M I m(t)(cid:3) Γ + 1 Γ m(0) − M I . (44) κI κII = Γ n2+1 2 . (39) Using Eq. (44), we can reduce the kinetic equations (29) to a system of two coupled different equations: Also primary nucleation is enhanced inside compartment I relative to compartment II: νI νII = Γn1 . (40) dM I m(t) dt dcI a(t) dt = −2k+ M I = k1 M I m(t) cI a(t) , m(t)n1 + k2 M I m(t)n2(cid:2)M I m(0) − M I (45a) m(t)(cid:3) , (45b) 2. Analytical solution for full time course of monomer concentrations in compartments I and II We now construct analytical solutions for the monomer and aggregate mass concentrations that are valid for the entire duration of the aggregation reaction. In the previ- ous section, we have seen that for Γ (cid:29) 1 two timescales, where k+ = k+Γ/(Γ + 1) and k2 = k2(Γ + 1)/Γ. Con- veniently, Eqs. (45) are exactly the fundamental kinetic equations describing the dynamics of protein aggregation in a pure system, i.e., without compartment, but with effective rate parameters that depend on the degree of phase separation [5, 46]. Thus, we can adapting the re- sults in [46] to Eq. (45), we find the following solution for (cid:20) (cid:32) (cid:20) the time varying monomer concentration in compartment I: (cid:18) Γ (cid:19) Γ + 1 (cid:21)−θ eκIt , (46) = 1 + λ2 I 2κ2 I θ M I M I m(t) m(0) (cid:112) 2/[n2(n2 + 1)]. Using Eq. (44), we then ob- where θ = tain an expression for the aggregate mass concentration: (cid:20) (cid:21)−θ(cid:33) (cid:18) Γ (cid:19) Γ + 1 eκIt . (47) M I M I a(t) m(0) = Γ + 1 Γ 1 − 1 + λ2 I 2κ2 I θ Finally, the time course of the monomer concentration in compartment II is obtained using the relationship M I m (t). This yields: m(t) = ΓM II (cid:18) Γ (cid:19) Γ + 1 (cid:21)−θ eκIt . (48) M II M II m (t) m (0) = 1 + λ2 I 2κ2 I θ The accuracy of our analytical solutions Eqs. (46), (47) and (48) against numerical integration of Eqs. (3) (main text) is shown in Fig. 5. From the knowledge of the time varying monomer con- centration, Eq. (46), we can obtain an expression for the aggregate number concentration in compartment I using Eq. (45a) by simple differentiation of Eq. (46), a(t) = −1/[2k+ M I cI m(t)/dt. This yields the fol- (cid:20) lowing expression: (cid:21)−1 m(t)]dM I cI a(t) cI a(∞) = 1 + 2κ2 I θ λ2 I where cI a(∞) = κIθ 2k+ (cid:18) Γ + 1 (cid:19) (cid:18) Γ + 1 Γ e −κIt (cid:19) Γ , (49) (50) is the number concentration of aggregates at steady state. It is interesting to extract from Eq. (50) the key depen- dence of cI a(∞) on the parameters ξ and Γ: cI a(∞) = κ0θ 2k+ 2 Γ n2−1 n2+1 ξ (Γ + 1) . 2 (51) Note that the prefactor defines the homogeneous concen- tration in the absence of compartments, Thus, using Γ (cid:29) 1 we find the following scaling relation- ship for the steady-state number concentration of aggre- gates in compartment I: 12 n2+1 2 . a (ξΓ) cI a(∞) (cid:39) c0 (53) A similar scaling relationship can be derived also for the steady-state number concentration of aggregates in com- partment II as follows. We recall that the early-time dynamics of aggregation in compartment II is character- ized by a timescale 1/κII, which is much slower than the timescale of aggregation in compartment I, 1/κI. Thus, cII a can be considered to be still in the exponential growing phase even when the aggregate concentration in compart- ment I is equilibrating. Eventually, the assembly in com- partment II is arrested abruptly as soon as aggregation in compartment I is fully saturated, since no monomer is left in either compartment. Since the timescale for satu- ration of aggregation in compartment I is proportional to 1/κI (see Eq. (46)), we can estimate the concentration of aggregates in compartment II at the end of the reaction as: (cid:16) (cid:17) − n2+1 2 Γ , (54) where in the last step we used Eq. (39). Since Γ (cid:29) 1, the argument of the sinh function is much smaller than unity. Hence, we can simplify Eq. (54) by using a Taylor expansion of the sinh function to first order, sinh x = x+O(x3), yielding the following scaling relationship after extracting the ξ dependence of νII and κII: cII a (∞) (cid:39) c0 aw ξn1− n2+1 2 Γ − n2+1 2 , (55) n1−n2−1. Combining Eq. (53) where w = k1/(k2θ) (M tot m ) with (55), we obtain one of the key results of our paper, namely the scaling behavior of aggregate enrichment be- tween compartments I and II with Γ: cI a (∞) ∝ ξn2+1−n1Γn2+1 , a(∞) cII (56) tive degree of monomer characterized by ξ(cid:0) ¯φ(cid:1) is given in where the impact of compartment volume on the rela- 3. Scaling relationships for the aggregate number concentrations in compartments I and II cII a (∞) (cid:39) νII sinh(κII/κI) κII = νII κII sinh c0 a = κ0θ 2k+ . (52) Eq. (28). [1] B. Alberts, Molecular biology of the cell (Garland science, [2] T. P. Knowles, M. Vendruscolo, and C. M. Dobson, Na- 2017). ture reviews Molecular cell biology 15, 384 (2014). 13 [3] F. Chiti and C. M. Dobson, Annu. Rev. Biochem. 75, 333 (2006). [4] A. D. Gitler, P. Dhillon, and J. Shorter, "Neurodegen- erative disease: models, mechanisms, and a new hope," (2017). [5] T. C. Michaels, A. Sari´c, J. Habchi, S. Chia, G. Meisl, M. Vendruscolo, C. M. Dobson, and T. P. Knowles, An- nual review of physical chemistry (2018). [26] M. Ganassi, D. Mateju, I. Bigi, L. Mediani, I. Poser, H. O. Lee, S. J. Seguin, F. F. Morelli, J. Vinet, G. Leo, et al., Molecular cell 63, 796 (2016). [27] S. Alberti, D. Mateju, L. Mediani, and S. Carra, Fron- tiers in molecular neuroscience 10, 84 (2017). [28] S. Alberti and S. Carra, Journal of Molecular Biology (2018). [29] S. Jain, J. R. Wheeler, R. W. Walters, A. Agrawal, [6] W. Neupert and J. M. Herrmann, Annu. Rev. Biochem. A. Barsic, and R. Parker, Cell 164, 487 (2016). 76, 723 (2007). [30] S. Specht, S. B. Miller, A. Mogk, and B. Bukau, J Cell [7] N. Wiedemann and N. Pfanner, Annual review of bio- Biol 195, 617 (2011). chemistry 86, 685 (2017). [8] J. Dukanovic and D. Rapaport, Biochimica et Biophysica [31] S. Alberti and A. A. Hyman, BioEssays 38, 959 (2016). [32] Y. Shin and C. P. Brangwynne, Science 357, eaaf4382 Acta (BBA)-Biomembranes 1808, 971 (2011). (2017). [9] C. P. Brangwynne, C. R. Eckmann, D. S. Courson, A. Rybarska, C. Hoege, J. Gharakhani, F. Julicher, and A. A. Hyman, Science 324, 1729 (2009). [10] C. P. Brangwynne, J Cell Biol 203, 875 (2013). [11] S. Elbaum-Garfinkle, Y. Kim, K. Szczepaniak, C. C.-H. Chen, C. R. Eckmann, S. Myong, and C. P. Brangwynne, Proceedings of the National Academy of Sciences 112, 7189 (2015). [12] L. Zhu and C. P. Brangwynne, Current opinion in cell biology 34, 23 (2015). [13] S. F. Banani, H. O. Lee, A. A. Hyman, and M. K. Rosen, Nature reviews Molecular cell biology 18, 285 (2017). [14] A. A. Hyman, C. A. Weber, and F. Julicher, Annual review of cell and developmental biology 30, 39 (2014). [15] C. P. Brangwynne, P. Tompa, and R. V. Pappu, Nature Physics 11, 899 (2015). [16] R. Parker and U. Sheth, Molecular cell 25, 635 (2007). [17] A. Patel, H. O. Lee, L. Jawerth, S. Maharana, M. Jahnel, M. Y. Hein, S. Stoynov, J. Mahamid, S. Saha, T. M. Franzmann, A. Pozniakovski, I. Poser, N. Maghelli, L. A. Royer, M. Weigert, E. W. Myers, S. Grill, D. Drechsel, A. A. Hyman, and S. Alberti, Cell 162, 1066 (2015). [18] L. Malinovska, S. Kroschwald, and S. Alberti, Biochim- ica et Biophysica Acta (BBA) - Proteins and Proteomics 1834, 918 (2013), the emerging dynamic view of pro- teins:Protein plasticity in allostery,evolution and self- assembly. [19] A. Molliex, J. Temirov, J. Lee, M. Coughlin, A. P. Kana- garaj, H. J. Kim, T. Mittag, and J. P. Taylor, Cell 163, 123 (2015). [20] A. Hern´andez-Vega, M. Braun, L. Scharrel, M. Jahnel, S. Wegmann, B. T. Hyman, S. Alberti, S. Diez, and A. A. Hyman, Cell reports 20, 2304 (2017). [21] J. B. Woodruff, B. F. Gomes, P. O. Widlund, J. Ma- and A. A. Hyman, Cell 169, hamid, A. Honigmann, 1066 (2017). [22] D. Mateju, T. M. Franzmann, A. Patel, A. Kopach, E. E. Boczek, S. Maharana, H. O. Lee, S. Carra, A. A. Hyman, and S. Alberti, The EMBO journal , e201695957 (2017). [23] S. Saha, C. A. Weber, M. Nousch, O. Adame-Arana, C. Hoege, M. Y. Hein, E. Osborne-Nishimura, J. Ma- hamid, M. Jahnel, L. Jawerth, et al., Cell 166, 1572 (2016). [24] H. Zhang, S. Elbaum-Garfinkle, E. M. Langdon, N. Tay- lor, P. Occhipinti, A. A. Bridges, C. P. Brangwynne, and A. S. Gladfelter, Molecular cell 60, 220 (2015). [25] T. M. Franzmann, M. Jahnel, A. Pozniakovsky, J. Ma- hamid, A. S. Holehouse, E. Nuske, D. Richter, W. Baumeister, S. W. Grill, R. V. Pappu, et al., Science 359, eaao5654 (2018). [33] C. M. Dobson, Nature 426, 884 (2003). [34] H. A. Lashuel, D. Hartley, B. M. Petre, T. Walz, and P. T. Lansbury Jr, Nature 418, 291 (2002). [35] S. M. Catalano, E. C. Dodson, D. A. Henze, J. G. Joyce, G. A. Krafft, and G. G. Kinney, Current topics in medic- inal chemistry 6, 597 (2006). [36] I. Benilova, E. Karran, and B. De Strooper, Nature neu- roscience 15, 349 (2012). [37] S. Campioni, B. Mannini, M. Zampagni, A. Pensalfini, C. Parrini, E. Evangelisti, A. Relini, M. Stefani, C. M. Dobson, C. Cecchi, and F. Chiti, Nature chemical biol- ogy 6, 140 (2010). [38] T. Grousl, S. Ungelenk, S. Miller, C.-T. Ho, M. Khokh- rina, M. P. Mayer, B. Bukau, and A. Mogk, J Cell Biol , jcb (2018). [39] E. E. Boczek and S. Alberti, J Cell Biol 217, 1173 (2018). [40] E. E. Griffin, D. J. Odde, and G. Seydoux, Cell 146, 955 (2011). [41] S. I. Cohen, S. Linse, L. M. Luheshi, E. Hellstrand, D. A. White, L. Rajah, D. E. Otzen, M. Vendruscolo, C. M. Dobson, and T. P. Knowles, Proceedings of the National Academy of Sciences 110, 9758 (2013). [42] P.-G. de Gennes, The journal of chemical physics 55, 572 (1971). [43] M. Rubinstein, Physical review letters 59, 1946 (1987). [44] M. Sajfutdinow, W. M. Jacobs, A. Reinhardt, C. Schnei- and D. M. Smith, Proceedings of the National der, Academy of Sciences , 201806010 (2018). [45] T. C. Michaels and T. P. Knowles, American Journal of Physics 82, 476 (2014). [46] T. C. Michaels, S. I. Cohen, M. Vendruscolo, C. M. Dob- son, and T. P. Knowles, Physical review letters 116, 038101 (2016). [47] P. Arosio, T. C. Michaels, S. Linse, C. Mansson, C. Emanuelsson, J. Presto, J. Johansson, M. Vendrus- colo, C. M. Dobson, and T. P. Knowles, Nature commu- nications 7 (2016). [48] F. Kundel, L. Hong, B. Falcon, W. A. McEwan, T. C. Michaels, G. Meisl, N. Esteras, A. Y. Abramov, T. J. Knowles, M. Goedert, et al., ACS chemical neuroscience (2018). [49] L. Zhu, X.-J. Zhang, L.-Y. Wang, J.-M. Zhou, and S. Perrett, Journal of molecular biology 328, 235 (2003). [50] A. M. Ruschak and A. D. Miranker, Proceedings of the National Academy of Sciences 104, 12341 (2007). [51] G. Meisl, X. Yang, E. Hellstrand, B. Frohm, J. B. Kirkegaard, S. I. Cohen, C. M. Dobson, S. Linse, and T. P. Knowles, Proceedings of the National Academy of Sciences 111, 9384 (2014). 14 [52] D. Zwicker, M. Decker, S. Jaensch, A. A. Hyman, and F. Julicher, Proceedings of the National Academy of Sci- ences 111, E2636 (2014). [58] N. Franks and W. Lieb, Nature 274, 339 (1978). [59] J. T. G. Overbeek and M. Voorn, Journal of Cellular and Comparative Physiology 49, 7 (1957). [53] W. H. Stockmayer, The Journal of chemical physics 11, [60] W. Ostwald, Zeitschrift fur physikalische Chemie 22, 289 45 (1943). (1897). [54] T. S. Harmon, A. S. Holehouse, M. K. Rosen, and R. V. [61] I. Lifshitz and V. Slyozov, Journal of Physics and Chem- Pappu, Elife 6 (2017). istry of Solids 19, 35 (1961). [55] T. P. Knowles, D. A. White, A. R. Abate, J. J. Agresti, S. I. Cohen, R. A. Sperling, E. J. De Genst, C. M. Dobson, and D. A. Weitz, Proceedings of the National Academy of Sciences 108, 14746 (2011). [56] P. Arosio, M. Vendruscolo, C. M. Dobson, and T. P. Knowles, Trends in pharmacological sciences 35, 127 (2014). [57] K. H. Meyer, Transactions of the Faraday Society 33, 1062 (1937). [62] A. Bray, Advances in Physics 43, 357 (1994), http://www.tandfonline.com/doi/pdf/10.1080/00018739400101505. [63] P. J. Flory, The Journal of chemical physics 10, 51 (1942). [64] M. L. Huggins, The Journal of Physical Chemistry 46, 151 (1942).
1806.05109
2
1806
2019-04-02T12:37:02
Statistical mechanics of an elastically pinned membrane: Static profile and correlations
[ "physics.bio-ph", "cond-mat.soft", "cond-mat.stat-mech" ]
The relation between thermal fluctuations and the mechanical response of a free membrane has been explored in great detail, both theoretically and experimentally. However, understanding this relationship for membranes, locally pinned by proteins, is significantly more challenging. Given that the coupling of the membrane to the cell cytoskeleton, the extracellular matrix and to other internal structures is crucial for the regulation of a number of cellular processes, understanding the role of the pinning is of great interest. In this manuscript we consider a single protein (elastic spring of a finite rest length) pinning a membrane modelled in the Monge gauge. First, we determine the Green$'$s function for the system and complement this approach by the calculation of the mode coupling coefficients for the plane wave expansion, and the orthonormal fluctuation modes, in turn building a set of tools for numerical and analytic studies of a pinned membrane. Furthermore, we explore static correlations of the free and the pinned membrane, as well as the membrane shape, showing that all three are mutually interdependent and have an identical long-range behaviour characterised by the correlation length. Interestingly, the latter displays a non-monotonic behaviour as a function of membrane tension. Importantly, exploiting these relations allows for the experimental determination of the elastic parameters of the pinning. Last but not least, we calculate the interaction potential between two pinning sites and show that, even in the absence of the membrane deformation, the pinnings will be subject to an attractive force due to changes in membrane fluctuations.
physics.bio-ph
physics
Manuscript submitted to BiophysicalJournal Article Statistical mechanics of an elastically pinned membrane: Static profile and correlations Josip Augustin Janeš1,2, Henning Stumpf1, Daniel Schmidt1,3, Udo Seifert3, and Ana-Sunčana Smith1,2,* 9 1 0 2 r p A 2 ] h p - o i b . s c i s y h p [ 2 v 9 0 1 5 0 . 6 0 8 1 : v i X r a 1PULS Group, Institut für Theoretische Physik and Cluster of Excellence: Engineering of Advanced Materials, Friedrich Alexander Universität Erlangen-Nürnberg, 91052 Erlangen, Germany 2Institut Ruđer Bošković, 10000 Zagreb, Croatia 3II. Institut für Theoretische Physik, Universität Stuttgart, 70569 Stuttgart, Germany *Correspondence: [email protected] ABSTRACT The relation between thermal fluctuations and the mechanical response of a free membrane has been explored in great detail, both theoretically and experimentally. However, understanding this relationship for membranes, locally pinned by proteins, is significantly more challenging. Given that the coupling of the membrane to the cell cytoskeleton, the extracellular matrix and to other internal structures is crucial for the regulation of a number of cellular processes, understanding the role of the pinning is of great interest. In this manuscript we consider a single protein (elastic spring of a finite rest length) pinning a membrane modelled in the Monge gauge. First, we determine the Green's function for the system and complement this approach by the calculation of the mode coupling coefficients for the plane wave expansion, and the orthonormal fluctuation modes, in turn building a set of tools for numerical and analytic studies of a pinned membrane. Furthermore, we explore static correlations of the free and the pinned membrane, as well as the membrane shape, showing that all three are mutually interdependent and have an identical long-range behaviour characterised by the correlation length. Interestingly, the latter displays a non-monotonic behaviour as a function of membrane tension. Importantly, exploiting these relations allows for the experimental determination of the elastic parameters of the pinning. Last but not least, we calculate the interaction potential between two pinning sites and show that, even in the absence of the membrane deformation, the pinnings will be subject to an attractive force due to changes in membrane fluctuations. 1 INTRODUCTION Most living cells and a number of their internal organelles are bounded by membranes, which are composed primarily of phospholipids and proteins. The latter, in selected cases, are designed to interact with neighbouring structures thereby pinning the membrane. As such, protein complexes become spatially coordinated, which has important consequences for the structural integrity of cells. A typical instance of such pinning is found in red blood cells, where the plasma membrane couples to the underlying spectrin network (1), although in this case additional forces associated with the soft scaffold will play a role. Another example is the pinning of the membrane to stiffer scaffolds such as actin. This affects a number of cellular functions (2), as it allows for the transmission of force (3), for example, during cell adhesion. In this case, proteins such as integrins or cadherins on the plasma membrane associate into supramolecular ensembles, binding the membrane to the cytoskeleton in the cell interior and, simultaneously, to the extracellular matrix or another cell (4). Similarly, inside the cell, for example on the nuclear envelope, the cytoskeleton again couples to the external nuclear membrane by nesprins, while toward the interior, protein p58 serves as a membrane attachment site for the nuclear lamina by acting as a specific receptor for lamin B (5). All these couplings regulate the mechanical state of the cell, which in turn affects the cell motility, division rate, proliferation, mechanosensitivity, and a number of other processes (4). Hence, understanding the principles of protein-mediated interactions between membranes and the surrounding scaffolds is one of the key problems in mechanobiology. Modeling pinned membranes, be it the adhesion process (6 -- 8), in the context of the interactions with the cytoskeleton (9), or the nuclear envelope (10), requires defining the force response at the single pinning site. While different models have been used in the literature (11 -- 13), the linear relation, where the protein attachment is described by a harmonic spring of a finite rest length, seems to capture a number of biological situations (14 -- 16). In particular, such models have been used for more than two decades to study the interplay between the pinning sites and the forces induced by the cytoskeleton, with the assumption that the role of the membrane is merely to provide spatial coordination to the proteins. However, it is becoming Manuscript submitted to Biophysical Journal 1 Janeš et. al Figure 1: Mean shape (cid:104)u(r)(cid:105) (gray line) and the spatially dependent fluctuation amplitude (cid:104)v2(r)(cid:105) (gray shaded area) of a membrane residing in a harmonic potential of strength γ at h0 separation from a flat substrate and pinned by an elastic spring of rest length l0. more obvious that the membrane itself is not a simple spectator, but that it can act as a regulatory component (17, 18), since it also produces forces (19). Nonetheless, because the membrane is in principle very soft, the pinning will have appreciable effects on the membrane itself. Already in the early theoretical works, it was demonstrated that protein-mediated attachments of the membrane affect its shape and fluctuations (11, 20, 21), a fact that was used to identify binding sites in cells and vesicles (22 -- 24). Subsequent simulations and analytical modeling showed that the mean membrane shape and roughness depend non-trivially on the instantaneous bond density (15, 25 -- 31). Alternative approaches showed, furthermore, that pinnings which experience strong frictional coupling in the membrane introduce corrections to the membrane tension (32). Polymeric anchors, on the other hand were found responsible for the rescaling of the bending stiffness of the composite membrane in a mode-dependent fashion (33). Another useful strategy relied on finding appropriate approximations to homogenize the pinning sites. As a result, a family of effective potentials that predict static properties of fluctuations were suggested in different regimes of fluctuation strength (34 -- 37). Many studies showed that membrane fluctuations depend on the properties of the pinning itself, such as the pinning's length and mechanical stiffness (18, 21, 38 -- 43). However, efforts to understand this coupling theoretically are scarce (13, 15, 36, 37, 44, 45). The difficulty lies in the pinning-induced coupling of plane wave modes or spherical harmonics (46, 47), which are otherwise independent in free membranes. The need to circumvent these technical problems led to the development of several computational approaches, which used the conveniences of Fourier transforms and plane wave basis sets (8, 28, 48), and allowed for the numerical evaluation of mode-coupling effects (31), or alternative basis sets (12). Ultimately these extensive simulations pointed to interesting many-body effects, which however could be distinguished from two-body interactions only in very limited regimes. In this manuscript, we provide a full analysis of static properties of a membrane pinned by an elastic spring (Fig. 1). We first calculate the static Green's function for the pinned membrane (section III), which is the working horse of analytic calculations. Given that they were not previously reported in the literature, we also provide explicit expressions for the orthonormal modes (Appendix A), and the mode coupling amplitudes for the plane wave expansion (Appendix B), both of which may be particularly useful for numerical calculations and the development of simulations, and show that they yield equivalent description as the Green's function approach. We use the Green's function to provide a comprehensive description of static properties of a pinned membrane in the full parameter range (section IV), focusing on the membrane's mean shape, fluctuation amplitude and the two-point spatial correlation function. Besides recovering the limits known in the literature for tensionless membranes and rigid pinning, our analysis of the correlation length (section V) elucidates the interplay between the membrane rigidity and tension, the strength of the nonspecific potential and the pinning elasticity. In the final section VI, we calculate explicitly and then analyze in detail the interaction potential and the force between two pinning sites. METHODS 2 THEORETICAL SETUP The system (Fig. 1) consists of one flexible pinning site (harmonic spring of an elastic constant λ and rest length l0, placed at the lateral position r0) that confines fluctuations of a tensed membrane (bending rigidity κ, tension σ). The membrane resides in the minimum of a harmonic non-specific potential (strength γ) at a height h0 above the substrate, except near r0, where it could be displaced by the pinning. 2 Manuscript submitted to Biophysical Journal u(r)h0γu2(r)l0rr0<u(r)><v2(r)> Static properties of a pinned membrane The membrane shape is parametrized in the linearized Monge gauge (49), such that u(r) denotes deviations from the shape of a flat membrane positioned in the minimum of the nonspecific potential along the lateral position r. Since pinnings typically introduce membrane displacements from the minimum (order of magnitude of 1-10 nm) (18, 50) that are small in comparison with the correlation length of the membrane (order of 100 nm), we use the linearized Hamiltonian (cid:35) (∇u(r))2 + + σ 2 (u(r))2 + γ 2 1 2 λ (u(r) − (l0 − h0))2 δ(r − r0) , (1) (cid:34) (cid:16)∇2u(r)(cid:17)2 H = dr κ 2 Þ A to describe the system. The first two terms in the integral on the right hand side comprise the Helfrich-Hamiltonian (51) for a bendable, pre-tensed membrane which resides in a nonspecific potential (third term). The energetic contribution of a harmonic spring for the pinning is represented by the fourth term which includes a delta function δ(r)) positioning the pinning, as further discussed in Supplementary Information (SI) section I. The integration goes over the projected membrane surface A. Here, and throughout the paper, the energy scale kBT (with Boltzmann constant kB and absolute temperature T), is set to unity. The validity of this Hamiltonian has been recently discussed in detail (52), where a reasonable agreement between numerically calculated and experimentally measured correlations and shapes has been obtained. With u(r) = (cid:104)u(r)(cid:105) + v(r), minimization of the Hamiltonian (eq. 1) provides the equation for the mean shape (cid:104)u(r)(cid:105) (cid:34) (cid:35) κ∇4 − σ∇2 + γ + λδ(r − r0) (cid:104)u(r)(cid:105) = λ(l0 − h0)δ(r − r0). (2) The fluctuations v(r) can be obtained from diagonalizing the second variation of the Hamiltonian, which leads to the eigenequation (cid:2)κ∇4 − σ∇2 + γ + λδ(r − r0)(cid:3) ψi(r) = Eiψi(r), (3) the latter containing the same operator as the shape equation 2. By expanding the fluctuations in these eigenmodes (see Appendix A) v(r) = aiψi(r), and using the equipartition theorem i we find the spatial two point correlation function (cid:104)aiaj(cid:105) = kBT Ei δi j (cid:104)v(r)v(r(cid:48))(cid:105) = i (r(cid:48)) ψi(r)ψ∗ Ei . i (4) (5) (6) We assume that the probability for membrane fluctuations with an amplitude larger than h0 is small, such that these configurations will not contribute significantly to the average properties of the membrane profile. With this assumption, the details of these configurations, which would involve a non-permeable boundary at the substrate, are not important, and we can instead deal with a simpler problem in which the substrate is completely permeable to the membrane. This approximation is satisfied if the protein that pins the membrane has a finite size (larger than the fluctuation amplitude of the membrane but smaller than h0), which is in experimental systems satisfied by the self-adjustment of the effective non-specific potential. Namely, if the proteins or the fluctuation amplitude of the membrane were larger than h0, this would renormalize the non-specific potential and move the minimum away from the substrate (hence h0 would be increased, and the curvature of the minimum, in our model captured by γ, would be changed), such that the required condition is recovered prior to the pinning. Practically, in the calculations this assumption is implied by having no boundary conditions on the amplitude of the membrane fluctuations. 3 GREEN'S FUNCTION APPROACH 3.1 Green's function for the free membrane Prior to addressing the problem of a pinned membrane, it is instructive to notice that the Green's function gf (rr(cid:48)) for the free membrane (λ = 0) is defined by (cid:2)κ∇4 − σ∇2 + γ(cid:3) gf (rr(cid:48)) = δ(r − r(cid:48)). (7) Manuscript submitted to Biophysical Journal 3 Janeš et. al It is translationally invariant (gf (rr(cid:48)) = gf (r − r(cid:48))) and can be expressed as Þ gf (r − r(cid:48)) = 1 (2π)2 dk eik(r−r(cid:48)) κk4 + σk2 + γ . (8) R2 The solution of the integral on the right hand side of eq. 8 is given in (53) and is a combination of modified Bessel functions of the second kind K0 gf (r − r(cid:48)) = K0(a−r − r(cid:48)) − K0(a+r − r(cid:48)) (cid:114) 1 −(cid:16) λ0 m 4σ (cid:17)2 2πσ Here, and the coefficients a± are given in the form a± = 1 ξ0 with 4σ λ0 m √ λ0 m = 8 κγ (cid:115) (cid:169)(cid:173)(cid:171)1 ± ξ0 = 4(cid:112) 1 − κ/γ. (cid:18) λ0 m 4σ 1/2 (cid:19)2(cid:170)(cid:174)(cid:172) , . (9) (10) (11) (12) We note that the Green's function eq. 9 is real even if a± are complex numbers. correlation function (cid:104)v f (r)v f (r(cid:48))(cid:105) and the mean square fluctuation amplitude (cid:104)v2 by several groups (38, 54 -- 56). The later is commonly denoted by 1/λm (8, 18, 44). Hence, As for any quadratic integral kernel, the Green's function gf (r − r(cid:48)), and respectively gf (0) are associated with the spatial f (r)(cid:105) of the free membrane, initially calculated (13) (14) (15) (16) (17) (cid:33) (cid:32)(cid:114)(cid:16) λ0 (cid:17)2 − 1 (cid:114)(cid:16) λ0 (cid:17)2 − 1 m 4σ m 4σ , gf (0) = 1 λm = arctan 2πσ which for a tensionless case (57) simplifies to gf (0)σ=0 = 1 λ0 m . Under this conditions, eq. 9 adopts the well-known form (21, 56) gf (r − r(cid:48))σ=0 = − 4 πλ0 m kei0 (cid:18) r − r(cid:48) (cid:19) ξ0 , with kei0 being the Kelvin function and ξ0 being the lateral correlation length of the free tensionless membrane given by eq. 12. 3.2 Green's function for the pinned membrane The Green's function g(rr(cid:48)) providing the response of a membrane at the position r due to a disturbance at the position r(cid:48) is defined as (cid:2)κ∇4 − σ∇2 + γ + λδ(r − r0)(cid:3) g(rr(cid:48)) = δ(r − r(cid:48)). (cid:2)κ∇4 − σ∇2 + γ(cid:3)(cid:2)g(rr(cid:48)) + λgf (rr0)g(r0r(cid:48)) − gf (rr(cid:48))(cid:3) = 0, With the use of eq. 7, eq. 16 can be recast as 4 Manuscript submitted to Biophysical Journal Static properties of a pinned membrane which can be generally valid only if the second bracket identically vanishes. Consequently, Setting r=r0 in eq. 18 provides g(rr(cid:48)) = gf (rr(cid:48)) − λgf (rr0)g(r0r(cid:48)). g(r0r(cid:48)) = gf (r0r(cid:48)) 1 + λgf (r0r0) = λm λ + λm gf (r0r(cid:48)), which, upon reinsertion into eq. 18, gives rise to the Green's function for the pinned membrane Although g(rr(cid:48)) is comprised of the translationally invariant gf (r − r(cid:48)), it itself is not generally translationally invariant. g(rr(cid:48)) = gf (r − r(cid:48)) − λλm λ + λm gf (r − r0)gf (r0 − r(cid:48)). 3.3 Representing shape and fluctuations By construction, g(rr0) differs only by a prefactor from the solution of the shape equation 2 (cid:104)u(r)(cid:105) = λ(l0 − h0)g(rr0). Combining eqs. 19 and 21 gives the mean shape As shown previously (15, 44), in the tensionless case combining eqs. 15 and 22 yields (cid:104)u(r)(cid:105) = λλm λ + λm (l0 − h0)gf (r − r0). (cid:104)u(r)(cid:105)σ=0 = − 4 π λ λ + λ0 m (l0 − h0)kei0 (cid:18) r − r0 (cid:19) ξ0 , (18) (19) (20) (21) (22) (23) (24) (25) (26) (27) which is a function of the kei function, as expected for the differential operator of the shape equation that is bilaplacian plus a constant (58, 59). In the limit of an infinitely stiff pinning λ → ∞, eq. 23 reproduces the result obtained in (21) . By comparing the bilinear expansion of the Green's function in the eigenfunctions ψj (eq. 3) g(rr(cid:48)) = j(r(cid:48)) ψj(r)ψ∗ Ej , and eq. 6, we find j (cid:104)v(r)v(r(cid:48))(cid:105) = g(rr(cid:48)), where the factor kBT = 1 on the right hand side is implicit. Hence, (cid:104)v(r)v(r(cid:48))(cid:105) = gf (r − r(cid:48)) − λλm λ + λm Naturally, by setting r(cid:48) = r in eq. 26 we obtain the fluctuation amplitude gf (r − r0)gf (r0 − r(cid:48)). with (cid:104)v2(r)(cid:105) = 1 λm − λλm λ + λm f (r − r0), g2 (28) The same result can be obtained by calculating the eigenfunctions ψj(r) for a system with a single pinning (see Appendix A.1) λ + λm . (cid:104)v2(r0)(cid:105) = 1 (cid:18) (cid:18)(cid:114) (cid:19)(cid:19)(cid:21) Jm(qr) + δm0Π(q) Ym(qr) + 2 π Km q2 + σ κ r , (29) 1 + δm0 (Π(q))2(cid:17) (cid:20) imeimφ (cid:114)(cid:16) ψm(r, q) = and using eq. 24 to obtain the Green's function (see Appendix A.2) . Manuscript submitted to Biophysical Journal 5 Janeš et. al Figure 2: Spatial dependence of the static properties of the membrane for varying pinning stiffness λ. a) The mean shape given by eq. 22 and b) the correlation function extracted from eq. 33 show the same properties. The overshoots in the shape coincide with anticorrelations presented in the insets. Note that a more conventional parametrization of the mean shape in terms of height above the substrate is trivially obtained with (cid:104)h(r)(cid:105) = h0 + (cid:104)u(r)(cid:105). Parameters: κ = 20kBT, σ = 10−20kBT/nm2, γ = 3 × 10−7kBT/nm4, and h0 − l0 = 1 nm. RESULTS 4 PROPERTIES OF THE MEAN SHAPE AND THE CORRELATION FUNCTION While the previous sections reveal the formal framework describing the effect of the pinning on the fluctuations of the membrane, several results warrant further discussion. Specifically, inserting the solution for the mean shape eq. 22 into the Hamiltonian eq. 1 determines the total elastic energy of the average configuration of the system (pinning and membrane) H [(cid:104)u(r)(cid:105)] = 1 2 λλm λ + λm (h0 − l0)2 ≡ 1 2 K(h0 − l0)2 (30) Equation 30 shows that the deformation energy increases quadratically with the height separation between the free membrane and the pinning, while it vanishes for h0 = l0 as described previously (15, 60). The prefactor K is an effective spring constant made up of two "springs" (the membrane and the pinning) connected in series, with λm being the membrane spring constant. From this point of view, K can be seen as the effective elastic constant of the system (8, 18, 44). The quadratic nature of eq. 30 is consistent with the quadratic form of the Hamiltonian eq. 1 and the "local" nature of the pinning. A further consequence is the linear relation between the mean shape and the correlation function from the pinning site (31) which emerges by inspection of eqs. 21 and 25. Here, the spatially independent prefactor has a form of a force on a harmonic spring. As a result, both the shape and the correlation function have the same features but due to a minus sign on the left hand side of eq. 31, the trends are opposite. For instance, the well-documented overshoot of the membrane shape (15, 21, 59) at distances of a couple of correlation lengths from the pinning is reflected in the anticorrelations in the same range (Fig. 2). Likewise, the displacement of the mean shape from the minimum of the non-specific potential increases with the increased pinning stiffness λ (Fig. 2a), while the amplitude of the pinning site correlation (cid:104)v(r)v(r0)(cid:105) decreases (Fig. 2b). (cid:104)u(r)(cid:105) = −λ(h0 − l0)(cid:104)v(r)v(r0)(cid:105), Interestingly, following eqs. 19 and 22, the correlation function and the mean shape can also be expressed in terms of the correlation function for the free membrane which emerges from the proportionality between the pinned- and the free-membrane correlation functions (cid:104)v(r)v(r0)(cid:105) (cid:104)v f (r)v f (r0)(cid:105) = (cid:104)v2(r0)(cid:105) (cid:104)v2 (r0)(cid:105) = f (cid:104)u(r)(cid:105) = −K(h0 − l0)(cid:104)v f (r)v f (r0)(cid:105), λm λ + λm . (32) (33) 6 Manuscript submitted to Biophysical Journal 0110∞048-1-0.50r-r0/ξ0〈u(r)〉/(h0-l0)468-0.01500.015λ/λm0a)0110∞04800.51r-r0/ξ0〈v(r)v(r0)〉/〈vf2(r0)〉468-0.01500.015λ/λm0b) Static properties of a pinned membrane Figure 3: Effect of the pinning on the membrane fluctuations (eq. 33). a) Varying σ and λ. b) Varying λ and λ0 m. This result clearly captures the interplay between the pinning stiffness λ and the parameters of the membrane (σ and λ0 m) which are combined in λm. If λ (cid:28) λm, the pinning does not affect membrane fluctuations, whereas if λ (cid:29) λm fluctuations at the pinning are completely suppressed and small changes in λ do not affect the system behavior. However, in the regime λ ≈ λm fluctuations can change noticeably, even for small change in the pinning stiffness (Fig. 3). Low tensed membranes will show such sensitivity if λ ≈ λ0 m (small λ/σ in Fig. 3b). Moreover, since the decay of correlations from the pinning site is independent of h0 and l0 (i. e., from the mean deformation), elastic properties of the pinning can be extracted directly from the change in the fluctuation amplitude between the pinned and the free states of the membrane. m (large λ/σ in Fig. 3b), while highly tensed membranes do so if λ (cid:29) λ0 Another interesting relation is the one between the spatially-dependent mean square fluctuation amplitude and the square of the membrane shape (cid:104)v(r)2(cid:105) = − 1 λm (cid:104)u(r)(cid:105)2 K(h0 − l0)2 . (34) Both of these features can be measured using reflection interference contrast microscopy (RICM) with very high accuracy (61). Using very sparsely distributed pinnings, and allowing for independent measurements of λm and h0 − l0, stiffness of the pinning becomes the only unknown parameter, which can thus be extracted by comparing the shape and fluctuation profiles. So far the stiffness of the proteins was typically measured using atomic force microscopy, but outside of membrane environment, so this relation opens a possibility to extract mechanical properties of the pinning protein in its native environment. Actually, the existence of such a relation has been inferred in imaging of pinning sites using RICM (22, 24). In these studies, the suppression of membrane fluctuations was used to identify pinning sites that are of a lateral dimension smaller than the optical resolution of the microscope, which was possible because the correlation length of the membrane was similar or larger than the diffraction limit of the setup. Further development of this approach relies however on the understanding of the dependence of the correlation length of the pinned membrane on system parameters, as provided herein. Manuscript submitted to Biophysical Journal 7 011010510-210-110010110210300.51σ/λm0〈v(r)v(r0)〉/〈vf(r)vf(r0)〉λ/λm0a)10-210010210410-210010210400.51λ/σ〈v(r)v(r0)〉/〈vf(r)vf(r0)〉λm0/σb) Janeš et. al 5 EFFECT OF THE MEMBRANE TENSION ON THE LONG-RANGE BEHAVIOR OF THE SHAPE AND CORRELATION FUNCTION Both the mean shape and the correlations from the pinning site are proportional to the free membrane correlations. Hence, the decay length of the correlation function will be that of the free membrane correlation function, implying the insensitivity of the correlation length and the deformation range to the length and stiffness of the pinning. Accordingly, dependent on various regimes (see SI section V for details), a power law and an oscillatory behavior are dominated by an exponential decay of a length ξ(κ, σ, γ) = ξ(ξ0, σ/λ0 m) identified through where = ξ ξ0 2 √ (cid:32) cos (cid:32) 1 2 arctan (cid:104)v f (r)v f (0)(cid:105) = gf (r) r→∞∼ e−r/ξ(ξ0,σ/λ0 m), (cid:33)(cid:33)(cid:33)−1 m 4σ (cid:32)(cid:114)(cid:16) λ0 (cid:17)2 − 1 (cid:17)2(cid:33)(cid:35)−1/2 if σ = 0, if 0 < σ < λ0 4 , m if σ = λ0 4 , m if σ > λ0 4 . m (cid:32) 1 (cid:34) 4σ λ0 m 1 − (cid:114) 1 −(cid:16) λ0 m 4σ  (35) (36) 4 λ0 Remarkably, increasing tension does not necessarily induce longer range height correlations. Instead, when bending dominates, small amounts of tension (σ < 1 m) actually reduce the decay length of correlations (Fig. 4). In this regime, the membrane shape and correlation function exhibit an overshoot / anti-correlations of the long range limit immediately after the pinning (Fig. 2), followed by an oscillatory behavior within an exponentially decaying envelope (Eq. SI-V.8). Similarly to systems that are governed only by bending and tension (no non-specific potential), the tension here flattens the membrane so that the spatial correlations decrease, due to changes in curvature which decay faster as the distance from the inclusion increases. m/4, the amplitude of the oscillations decreases. When Specifically, as the tension increases toward the critical value of σc = λ0 the tension reaches σc, the oscillations are completely flattened, and the system enters a tension dominated regime. Now, coupling to the non-specific potential induces a slow, purely exponential decay of the shape and the correlations (SI-V.7). In this case, the larger the tension, the longer the range of the deformation and the correlation function, simply because of the increase m/16 the in the energy penalty for large curvatures in a nonspecific potential. However, only when the tension reaches σ = 5λ0 correlation length becomes longer than that of a tensionless free membrane. Notably, the mean shape and correlations (and their derivatives with respect to the spatial coordinate r) are continuous functions of σ, even at σc, and no actual singularity appears in the system at the crossover between the bending and tension dominated regimes. 6 MEMBRANE-MEDIATED INTERACTIONS BETWEEN TWO PINNINGS Equations 35 and 36 are significant in the context of interactions between pinnings on the membrane separated by a relative distance x. Following previous work (15) the interaction energy between two pinnings is V2(x) = K(l0 − h0)2 1 + Kgf (x) + 1 2 ln (cid:16) 1 −(cid:2)Kgf (x)(cid:3)2(cid:17) , (37) where the first term is the deformation energy stored in the system with two bonds and the second term is the entropic cost associated with the suppression of fluctuations (see SI Section IV for details of the calculation). Terms which are independent on the relative distance between the two pinnings are omitted, since they drop out in the calculation of the force between two pinnings F2(x) = −∂V2(x)/∂x, which becomes F2(x) = K2(l0 − h0)2g(cid:48) f (x) [1 + Kgf (x)]2 + K2gf (x)g(cid:48) f (x) 1 − K2gf (x)2 . (38) 8 Manuscript submitted to Biophysical Journal Static properties of a pinned membrane Figure 4: Correlation decay length ξ(ξ0, σ/λ0 σ = (5/16)λ0 m, beyond which an increase in tension results in longer range correlations than in the tensionless case. m) in the asymptotic limit r → ∞ (eq. (36)). The dark grey lines mark the value Thus, the spatial dependence of the force is given by the correlation function of a free membrane at the relative distance x. The first term on the right hand side of eq. 38 can be associated with the force that emerges due to the membrane deformation, while the second term is the force arising from the suppression of membrane fluctuations in a spatially-dependent manner. If the pinning deforms the membrane (h0 (cid:44) l0), the deformation term determines the long-range behaviour of the force, as it f (x), which decays decays two times slower than the fluctuation term (Fig. 5). Namely, the deformation term is proportional to g(cid:48) exponentially, and independent of the amount of the deformation in the system, while the fluctuation term, being proportional to gf (x)g(cid:48) f (x), decays exponentially but twice as fast (Fig. 5a). The deformation term typically dominates closer to the pinning as well (Fig. 5). However, if h0 (cid:39) l0, fluctuation forces dominate, in which case the decay length of the force is halved in comparison to the case of a deformed membrane. This means that even if the protein does not affect the membrane shape (h0 = l0), significant force may emerge and potentially lead to the agglomeration of pinning sites, as suggested by simulations of a membrane containing many pinnings, described by the same Hamiltonian (8, 31, 37, 48). While only limited understanding of the conditions necessary for the formation of domains is available at the moment, access to eq. 38 sets the foundation of the calculation of critical parameters which are necessary for the process of agglomeration. Based on the qualitative behaviour of the forces, we can recognize two regimes, namely the bending dominated (σ < σc) and the tension dominated regime (σ > σc). These regimes correspond to different regimes of the correlation function (see SI section V.). In the bending dominated regime, a repulsive barrier appears in the force at distances of few membrane correlation lengths (Fig. 5a). Increasing tension, but staying under σc, flattens the barrier and the oscillating tail of the force (Fig. 5a - inset). This is contrasted by the tension dominated regime in which the repulsive barrier and the oscillating tail disappear and the long range forces are attractive (Fig. 5b). Moreover, the range of the force increases with tension (Fig. 5b - inset). In all cases the range of these weak interactions is of the order of 100 nm which is nearly two orders of magnitude more than the direct protein-protein interactions. They are therefore considered long range, despite their universally-exponential nature. This exponential decay is contrasted by a body of work performed on forces between membrane inclusions in "bending only", or "tension only" systems for which the differential operator exhibits no scale. In the former case, the Green's function behaves as gf (r) ∼ r2 log r2 (62 -- 64), and switching tension affects the power law nature of the decay (40, 65, 66). Because the nonspecific potential introduces a length scale, the pinned membrane clearly delineates from these models for inclusions. However, it was recently proposed that a Hamiltonian, which is mathematically identical to that in eq. 1 can be used to model the inclusion of a protein with hydrophobic mismatch into a membrane (67). Although the parameter range in which the linearized theory is valid could be more narrow than in the case of pinnings, the analogy of formalisms between the two problems, in principle, allows for the exploitation of the current results. Consequently, exponential decays should also appear in forces acting between membrane inclusions. However, these forces will have very different magnitudes and overall range. It is worth mentioning that so far we neglected the finite size of proteins. This is appropriate for sparse or immobile protein attachments (size of the attachment is still smaller than the correlation length of the membrane). When proteins approach within a few nanometers separations between their surfaces, direct protein interactions will compete with the typically attractive membrane mediated interactions. The result of this competition at short range is non-universal, and is most likely dominated by Manuscript submitted to Biophysical Journal 9 1/45/161.01.21.41.61.8σ/λm0ξ/ξ0 Janeš et. al m < 1/4) is shown. For this specific set of parameters, Figure 5: Force between two pinnings. a) Bending dominated regime (σ/λ0 the deformation and the fluctuation contributions to the force are comparable. Both contributions oscillate around zero (inset), but the fluctuation part decays two times faster. b) The tension dominated regime (σ/λ0 m > 1/4). As we increase the tension, the range of the force increases. Parameters: λ = 0.75 × 10−2kBT/nm2, κ = 20kBT, γ = 3.125 × 10−7kBT/nm4, h0 − l0 = 10 nm. the direct contributions. Our hope is that the current approaches can be expanded to account for this case - either using the GF approach in analytic calculations, or using the expansions into relevant basis set for numerical simulations. 7 DISCUSSION AND CONCLUSIONS In this paper, we studied the effect of a pinning on the statics of a membrane fluctuating in a harmonic nonspecific potential. We showed that the membrane and the pinning can be seen as two springs in series in the context of the energetics, as discussed previously (18). Hence, in the case when the length of the pinning does not coincide with the position of the undisturbed membrane in an effective potential, the deformation in the system depends on the effective spring constants of the pinning and the membrane (the later characterized by the inverse of the fluctuation amplitude in the absence of pinning). For stiff membranes, the pinning will extend its shape, while for stiff pinnings, membrane deformation will be considerable. However, since the lateral correlation length of the membrane is not affected by the pinning properties, the range of the deformation is independent of the pinning. This is very different to the effect of tension, which directly affects the correlation length, in a non-homogeneous fashion. The pinning, on the other hand, has a major effect on the membrane fluctuation amplitude, which is an inverse function of the pinning stiffness. The correlation length and the long range exponential behavior is, however, fully given by the correlation length of the free membrane. For small tensions, a pinning may induce short-range anticorrelations of fluctuations and an overshoot of the membrane shape. In this regime, the correlation length decreases with increasing tension. At high tensions, the correlation length increases, while the shape and the correlations continuously decay to their long-range limits. These correlations translate into long-range interactions between pinnings, which also decay exponentially. The forces associated with this interaction potential are stronger if the pinning displaced the membrane, however, even in the absence of the deformation, the pinnings interact due to the suppression of fluctuations, analogously to Casimir forces. The results presented here open the possibility for differentiating between actively and passively pinned membranes in experiments, just by measuring the shape and fluctuations around a binding site, which can be either a single protein or a nanodomain, when the line tension remains small. Violation of the relationship (eq. 31 - 34) between the correlation functions and the shape provided in Section IV could be taken as a notion of activity. Moreover, in passive systems with small non-linear effects, exploiting the same relations could provide the foundation for the measurement of the stiffness of proteins in their natural membrane environment. The here-proposed models should be suitable for analysis of data obtained using interferometric methods, or in conjunction with atomic force microscopy of membrane-protein interactions, where vesicles are used as soft probes. Given that membranes, locally pinned by proteins or macromolecular assemblies, are indeed ubiquitous in nature, a toolbox developed herein consisting of mode-coupling coefficients, orthonormal modes and the Green's function of the system is highly 10 Manuscript submitted to Biophysical Journal 040010000-4-8r-r0[nm]ℱ[10-4kBTnm-1]4001000160000.1σ/λm0=0σ/λm0=1/20σ/λm0=1/5ℱ2eq.(37)1sttermeq.(37)2ndtermeq.(37)a)400800-50-2r-r0[nm]ℱ2[10-4kBTnm-1]4001000-0.30-0.1σ/λm0=1/3σ/λm0=1σ/λm0=3b) Static properties of a pinned membrane useful for future theoretical studies of membranes which aim to elucidate the interplay between the membrane elasticity and the forces transmitted by the proteins in the biological context. We may anticipate that the Green's function approach may be the method of choice for analytic modeling, however, normal modes and the mode coupling coefficients for the plane waves may be particularly useful in the context of numerical calculations. Of course, the equivalence of all three approaches can be stated by construction. Nevertheless, in terms of results presented herein, GF and plane wave approaches give exactly the same representation of the mean shape (eq. 22 vs eq. 70) and the correlations (eq. 26 vs eq. 71), while the normal modes give an alternative, but numerically identical representation (eq. 58 and eq. 60 for the mean shape and correlations, respectively). Besides studies in which membranes are used as probes for proteins binding during cell-cell and cell-substrate adhesion, or in the analysis of the interaction of the cytoskeleton with the plasma or nuclear membranes, which were in some cases based on the same Hamiltonian, other systems may benefit from the here developed tools and relations. In particular, as pointed out in the recent work of (67), the same Hamiltonian could be used in studies of the interactions between membrane inclusions (49, 68, 69). However, since the energetics and the length scales of characteristic interactions are very different, non-linear corrections may become important. As there is a wealth of systems where protein mediated pinning is important in the biological and biotechnological context, further developing a theory to account for the fluctuation dynamics of a permanently, but also stochastically pinned membranes appears as a natural and necessary extension of the current work, a task that we plan to undertake in our future work. Acknowledgments: A.-S.S. D.S. H.S and J.A.J were funded by ERC Starting Grant MembranesAct 337283. J.A.J and A.-S.S were in part supported by Croatian Science Foundation research project CompSoLs MolFlex 8238. D.S. was member of the Research Training Group 1962 at the Friedrich-Alexander-Universität Erlangen-Nürnberg. Author Contributions: A.S.S and U.S conceived the study. A.S.S was in charge of overall direction and supervision. U.S.provided critical feedback and helped shape the research. D.S. obtained the mode-coupling coefficients. J.A.J. and D.S calculated the orthonormal modes. J.A.J. developed the Green's function approach. H.S. performed the asymptotic analysis. The force between two pinnings was calculated by H.S. and J.A.J.. A.S.S, J.A.J., and H.S. wrote the manuscript with contributions from all authors. The authors declare that they have no competing interests. All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials. All data and computer code for this study are available on request from the authors. A NORMAL MODES EXPANSION A.1 Solution of the eigenmode equation It remains to determine the normal modes ψj given by eq. 3. By placing the pinning at the origin (r0 = 0), the solution of eq. 3 obeys radial symmetry with respect to the pinning site. Hence, the eigenmodes are a product of axial and radial functions, characterized by relevant mode numbers m and n, respectively where (r, φ) are polar coordinates of the position r. In this case, eq. 3 takes the form ψnm(r) = Rnm(r)eimφ, (cid:2)κ∇4 − σ∇2 + γ + λδ(r)(cid:3) ψnm(r) = Enmψnm(r), (39) (40) where Enm are the eigenvalues corresponding to modes {n, m}. The square brackets on the left hand side enclose the energy operator which must be Hermitian (SI section II.A.1). The general solution of eq. 40 emerges as a sum of Bessel functions (SI section II.B) Rnm(r) =anmJm(qnmr) + bnmYm(qnmr) + cnmKm(Qnmr) + dnmIm(Qnmr) with (cid:114) Qnm = q2 nm + . σ κ (41) (42) Here, Jm and Ym are Bessel functions of the first and second kind, Km and Im are the modified Bessel functions of the first and second kind, respectively, and anm, bnm, cnm and dnm are coefficients associated with the n and m mode numbers. The corresponding eigenvalues in eq. 40 are given by Enm = κq4 nm + σq2 nm + γ, (43) Manuscript submitted to Biophysical Journal 11 Janeš et. al and the general solution Rnm(r) is specified by appropriate boundary conditions. Boundary condition 1 - Rnm(r) stays finite when r → 0: The Bessel functions of the first kind, Jm and Im, inherently fulfill this boundary condition (J0(0) = I0(0) = 1 and Jm(0) = Im(0) = 0 for m > 0). The remaining Bessel functions Ym and Km diverge for r → 0. However, for m = 0, both Bessel functions diverge logarithmically such that the sum b0Y0(qnmr)+c0K0(Qnmr) stays finite with c0 = 2b0/π, while for m > 0 such cancellation is not possible. Consequently, (cid:18) (cid:19) Rnm(r) =anmJm(qnmr) + dnmIm(Qnmr) + δm0bnm Ym(qnmr) + Km(Qnmr) 2 π , (44) where δm0 is the Kronecker delta. The term multiplied by δm0 is contributing only for m = 0. Boundary condition 2 - The integral of the eigenvalue equation 40 over an infinitesimally small disk D() centered at the pinning has to vanish, Þ dr(cid:2)κ∇4 − σ∇2 + γ + λδ(r)(cid:3) ψnm(r) = 0. D() This boundary condition, often introduced around a delta function, is necessary to ensure the finiteness of the membrane profile at the origin. With this imposed, the integration of the right hand side of the eigenequation 5 vanishes in the relevant limit, and the limit is well defined. By extension, the integral of the left-hand side of the eigenequation 5 vanishes too (see SI section II. for details). By solving the integral for each mode, one obtains where bn0 = Π(qn0)(an0 + dn0), π ln(1 + σ κq2 n0 Boundary conditions 3 and 4 - At the membrane edge, r = P, we have 2κ) + λ 8κ(q2 n0 + σ Π(qn0) = ) . (46) (47) (48) (49) Rnm(P) = 0 ∆Rnm(P) = 0, λ (cid:18) (45) (50) (51) (52) in the eigenvalue equation 40 as shown in SI section II. where ∆ denotes the Laplacian operator. These boundary conditions arise in pair after imposing hermiticity of the operator From eqs. 46-49 we obtain the asymptotic form of Rnm(r) for a large membrane radius P (SI section II.C.2) (cid:26) (cid:20) Jm(qnmr) +δm0 Π(qn0) Ym(qnmr) + Km(Qnmr) 2 π (cid:19)(cid:21)(cid:27) , Rnm(r) ∼ anm with This asymptotic form of qnm emerges when n → ∞ and membrane radius P → ∞ as shown in SI-Section II.B-C. Normalization of the solution of the eigenvalue problem (SI section II.C.2), requires setting Finally, by letting P → ∞, qnm → q ∈ R, the basis functions become ψm(r, q), and are given by (SI section III.C.3) ψm(r, q) = Jm(qr) + δm0Π(q) Ym(qr) + 2 π Km q2 + σ κ r . (53) (cid:18)(cid:114) (cid:19)(cid:19)(cid:21) qnm ∼ n π P . (cid:114)(cid:16) im 1 + δm0 (Π(n∆q))2(cid:17) . (cid:18) anm = 1 + δm0 (Π(q))2(cid:17) (cid:20) imeimφ (cid:114)(cid:16) 12 Manuscript submitted to Biophysical Journal Naturally, the orthogonality condition Þ R2 Static properties of a pinned membrane drψm(r, q)ψ m(cid:48)(r, q(cid:48)) = ∗ δ(q − q(cid:48)) q 2πδm,m(cid:48) (54) is satisfied, and the profile of an infinite pinned membrane can be expanded in the basis functions ψm(r, q) as (SI section II.C.4) with where u(r) = 1 2π dqqUm(q)ψm(r, q), m=−∞ 0 Þ ∞ ∞ Þ ∞ Þ 2π 0 Um(q) =2π drr um(r)R∗ m(r, q), um(r) = 1 2π 0 dφ u(r)e−imφ. (55) (56) (57) (58) (59) (60) (61) (62) = 1 λ + λm , For vanishing λ (SI section II.C.5) the eigenmodes are given by the Bessel functions Jm(qr) for all m, which is equivalent to a basis set constructed from plane waves in radial geometry, as demonstrated for a free membrane. For a non-vanishing λ, on the other hand, the pinning properties affect explicitly only the eigenmode with m = 0. A.2 Representing shape and fluctuations Expansion of the mean shape of the membrane pinned at r0 = 0 is given only by m = 0 modes (SI section II.D.2): (cid:104)u(r)(cid:105) = λ(l0 − h0) 1 2π 0(0, q) R0(r, q)R∗ Eq . dqq Þ ∞ 0 At the pinning site r = 0 (cid:104)u(0)(cid:105) =λ(l0 − h0) 1 2π The correlation function is given by Þ ∞ 0 (cid:104)v(r1)v ∗(r2)(cid:105) = g(r1r2) = and the fluctuation amplitude by ψ0(0, q)2 dqq = λ λ + λm (l0 − h0). dqq ψm(r1, q)ψ∗ Eq m(r2, q) , Eq Þ ∞ 0 ∞ Þ ∞ m=−∞ 1 2π ∞ m=−∞ 0 (cid:104)v2(r)(cid:105) = 1 2π dqq ψm(r, q)2 Eq . At the position of the pinning site ∞Þ 0 (cid:104)v2(0)(cid:105) = 1 2π dq q κq4 + σq2 + γ (cid:0)8κq2 + 4σ(cid:1)2 λ2 +(cid:2)8κq2 + 4σ + λ π ln(cid:0)1 + σ/(κq2)(cid:1)(cid:3)2 The last equality, which coincides with eq. 27, was checked numerically to the machine precision for an arbitrary tension, and analytically for σ = 0. Manuscript submitted to Biophysical Journal 13 Janeš et. al B PLANE WAVE EXPANSION B.1 Mode-coupling Relating the shape and the fluctuation amplitude to the properties of the free membrane should be also possible in the most commonly used plane wave expansion Þ Þ R2 R2 u(r) = 1 (2π)2 (cid:104)u(r)(cid:105) = 1 (2π)2 dk u(k)eikr, dk(cid:104)u(k)(cid:105)eikr, Þ Þ where for the mean shape we find and for the correlation function Þ Þ (63) (64) (66) (67) (68) (69) (70) (71) (72) 1 (2π)4 (cid:104)v(r)v(r(cid:48))(cid:105) = (65) The disadvantage of this approach is the coupling of the modes, giving rise to expansion coefficients (cid:104)u(k)u(k(cid:48))(cid:105) that have so far not been calculated explicitly. As previously discussed (15), the amplitudes (cid:104)u(k)(cid:105) and the mode coupling coefficients (cid:104)u(k)u(k(cid:48))(cid:105) are defined as dk(cid:48)(cid:104)u(k)u(k(cid:48))(cid:105)eikreik(cid:48)r(cid:48) − 1 (2π)4 dk(cid:48)(cid:104)u(k)(cid:105)(cid:104)u(k(cid:48))(cid:105)eikreik(cid:48)r(cid:48). dk R2 R2 dk R2 R2 (cid:104)u(k)(cid:105) ≡ 1 Z (cid:104)u(k)u(k(cid:48))(cid:105) ≡ 1 Z Þ Þ Þ D[u] u(k) exp[−H] , D[u] u(k)u(k(cid:48)) exp[−H] , D[u] exp[−H] . with Z being the partition function Z = Treating identities in eq. 66 as Gaussian integrals (SI section III), gives (cid:104)u(k)(cid:105) = − λλm λ + λm δ(k + k(cid:48)) (h0 − l0) (cid:104)u(k)u(k(cid:48))(cid:105) = κk4 + σk2 + γ e−ikr0 κk4 + σk2 + γ , + (cid:104)u(k)(cid:105)(cid:104)u(k(cid:48))(cid:105) − λλm λ + λm e−ikr0 κk4 + σk2 + γ e−ik(cid:48)r0 κk(cid:48)4 + σk(cid:48)2 + γ . B.2 Representing shape and fluctuations Combining eqs. (64) and (68) we obtain the mean shape for a pinned membrane (l0 − h0)gf (r − r0). (cid:104)u(r)(cid:105) = λλm λ + λm By combining eqs. (65), (68) and (69), we obtain for the spatial correlations and for the fluctuation amplitude (r = r(cid:48)) (cid:104)v(r)v(r(cid:48))(cid:105) = gf (r − r(cid:48)) − λλm λ + λm gf (r − r0)gf (r0 − r(cid:48)) (cid:104)v2(r)(cid:105) = 1 λm − λλm λ + λm f (r − r0). g2 We have therefore independently derived the same result as with the Green's function approach (eqs. 22 and 26). SUPPORTING CITATIONS References (70 -- 72) appear in the Supporting Material. 14 Manuscript submitted to Biophysical Journal Static properties of a pinned membrane REFERENCES 1. Gov, N., and S. Safran, 2004. Pinning of fluid membranes by periodic harmonic potentials. Phys. Rev. E 69:011101. 2. Sackmann, E., and A.-S. Smith, 2014. Physics of cell adhesion: some lessons from cell-mimetic systems. Soft Matter 10:1644 -- 1659. 3. Schwarz, U. S., and S. A. Safran, 2013. Physics of adherent cells. Rev. Mod. Phys. 85:1327 -- 1381. 4. Hu, X., F. M. Margadant, M. Yao, and M. P. Sheetz. Molecular stretching modulates mechanosensing pathways. Protein Science 26:1337 -- 1351. https://onlinelibrary.wiley.com/doi/abs/10.1002/pro.3188. 5. Worman, H. J., J. Yuan, G. Blobel, and S. D. Georgatos, 1988. A lamin B receptor in the nuclear envelope. Proceedings of the National Academy of Sciences 85:8531 -- 8534. http://www.pnas.org/content/85/22/8531. 6. Blokhuis, E. M., and W. F. C. Sager, 1999. Helfrich free energy for aggregation and adhesion. The Journal of Chemical Physics 110:3148 -- 3152. https://doi.org/10.1063/1.478190. 7. Erdmann, T., and U. S. Schwarz, 2006. Bistability of cell-matrix adhesions resulting from nonlinear receptor-ligand dynamics. Biophys. J. 91:L60 -- L62. 8. Bihr, T., U. Seifert, and A.-S. Smith, 2015. Multiscale approaches to protein-mediated interactions between mem- branes -- relating microscopic and macroscopic dynamics in radially growing adhesions. New J. Phys. 17:083016. 9. Alert, R., J. Casademunt, J. Brugués, and P. Sens, 2015. Model for Probing Membrane-Cortex Adhesion by Micropipette Aspiration and Fluctuation Spectroscopy. Biophys. J. 108:1878 -- 1886. 10. Lammerding, J., 2011. Mechanics of the Nucleus, American Cancer Society, 783 -- 807. 11. Menes, R., and S. A. Safran, 1997. Nonlinear response of membranes to pinning sites. Phys. Rev. E 56:2:1891 -- 1899. 12. Lin, L. C.-L., J. Groves, and F. L. H. Brown, 2006. Analysis of Shape, Fluctuations, and Dynamics in Intermembrane Junctions. Biophys. J. 91:3600 -- 3606. 13. Netz, R. R., 1997. Inclusions in Fluctuating Membranes: Exact Results. J. Phys. I France 7:833 -- 852. 14. Seifert, U., 2000. Rupture of multiple parallel molecular bonds under dynamic loading. Phys. Rev. Lett. 84:2750. 15. Schmidt, D., T. Bihr, U. Seifert, and A.-S. Smith, 2012. Coexistence of dilute and densely packed domains of ligand-receptor bonds in membrane adhesion. Europhys. Lett. 99:38003. 16. Bauer, M., P. Kékicheff, J. Iss, C. Fajolles, T. Charitat, J. Daillant, and C. M. Marques, 2015. Sliding tethered ligands add topological interactions to the toolbox of ligand -- receptor design. Nature Communications 6:8117. 17. Perez, T. D., M. Tamada, M. P. Sheetz, and W. J. Nelson, 2008. Immediate-Early Signaling Induced by E-cadherin Engagement and Adhesion. Journal of Biological Chemistry 283:5014 -- 5022. 18. Fenz, S., T. Bihr, D. Schmidt, R. Merkel, U. Seifert, K. Sengupta, and A.-S. Smith, 2017. Membrane fluctuations mediate lateral interactions between cadherin bonds. Nature Physics 13:906 -- 913. 19. Bell, G., 1978. Models for the specific adhesion of cells to cells. Science 200:618 -- 627. 20. Dan, N., P. Pincus, and S. A. Safran, 1993. Membrane-induced interactions between inclusions. Langmuir 9:2768 -- 2771. 21. Bruinsma, R., M. Goulian, and P. Pincus, 1994. Self-assambly of membrane junctions. Biophys. J. 67:746 -- 750. 22. Smith, A.-S., K. Sengupta, S. Goennenwein, U. Seifert, and E. Sackmann, 2008. Force-induced growth of adhesion domains is controlled by receptor mobility. Proc. Natl. Acad. Sci. U. S. A. 105:6906 -- 6911. 23. Pierres, A., A.-M. Benoliel, D. Touchard, and P. Bongrand, 2008. How Cells Tiptoe on Adhesive Surfaces before Sticking. Biophys. J. 94:4114. 24. Smith, A.-S., S. Fenz, and K. Sengupta, 2010. Inferring spatial organization of bonds within adhesion clusters by exploiting fluctuations of soft interfaces. Europhys. Lett. 89:28003:1 -- 6. Manuscript submitted to Biophysical Journal 15 Janeš et. al 25. Lin, L. C.-L., and F. L. H. Brown, 2004. Dynamics of pinned membranes with application to protein diffusion on the surface of red blood cells. Biophys. J. 86:764 -- 780. 26. Gov, N. S., and S. A. Safran, 2005. Red blood cell membrane fluctuations and shape controlled by ATP-induced cytoskeletal defects. Biophys. J. 88:1859 -- 1874. 27. Smith, A.-S., and U. Seifert, 2005. Effective adhesion strength of specifically bound vesicles. Phys. Rev. E 71:061902. 28. Lin, L. C.-L., N. Gov, and F. L. H. Brown, 2006. Nonequilibrium membrane fluctuations driven by active proteins. J. Chem. Phys. 124:074903. 29. Krobath, H., G. J. Schütz, R. Lipowsky, and T. R. Weikl, 2007. Lateral diffusion of receptor-ligand bonds in membrane adhesion zones: Effect of thermal membrane roughness. Europhys. Lett. 78:38003. 30. Reister, E., T. Bihr, U. Seifert, and A.-S. Smith, 2011. Two intertwined facets of adherent membranes: membrane roughness and correlations between ligand -- receptors bonds. New J. Phys. 13:025003:1 -- 15. 31. Fenz, S. F., T. Bihr, R. Merkel, U. Seifert, K. Sengupta, and A.-S. Smith, 2011. Switching from Ultraweak to Strong Adhesion. Adv. Mater. 23:2622 -- 2626. 32. Fournier, J.-B., D. Lacoste, and E. Raphaël, 2004. Fluctuation spectrum of fluid membranes coupled to an elastic meshwork: jump of the effective surface tension at the mesh size. Phys. Rev. Lett. 92:018102. 33. Auth, T., and G. Gompper, 2005. Fluctuation spectrum of membranes with anchored linear and star polymers. Phys. Rev. E 72:031904. 34. Breidenich, M., R. R. Netz, and R. Lipowsky, 2000. The shape of polymer-decorated membranes. Europhys. Lett. 49:431 -- 437. 35. Merath, R.-J., and U. Seifert, 2006. Nonmonotonic fluctuation spectra of membranes pinned or tethered discretely to a substrate. Phys. Rev. E 73:010401. 36. Farago, O., 2008. Membrane fluctuations near a plane rigid surface. Phys. Rev. E 78:051919. 37. Speck, T., E. Reister, and U. Seifert, 2010. Specific adhesion of membranes: Mapping to an effective bond lattice gas. Phys. Rev. E 82:021923. 38. Seifert, U., 1997. Configurations of fluid membranes and vesicles. Adv. Phys. 46:13 -- 137. 39. Weikl, T. R., 2001. Fluctuation-induced aggregation of rigid membrane inclusions. Europhys. Lett. 54:547 -- 553. 40. Evans, A. R., M. S. Turner, and P. Sens, 2003. Interactions between proteins bound to biomembranes. Phys. Rev. E 67:10. 41. Lin, L. C.-L., and F. L. H. Brown, 2006. Simulating Membrane Dynamics in Nonhomogeneous Hydrodynamic Environments. J. Chem. Theory Comput. 2:472 -- 483. 42. Hu, J., R. Lipowsky, and T. R. Weikl, 2013. Binding constants of membrane-anchored receptors and ligands depend strongly on the nanoscale roughness of membranes. Proc. Natl. Acad. Sci. U. S. A. 110:15283 -- 15288. 43. Dharan, N., and O. Farago, 2016. Formation of semi-dilute adhesion domains driven by weak elasticity-mediated interactions. Soft Matter 12:6649 -- 6655. 44. Bihr, T., U. Seifert, and A.-S. Smith, 2012. Nucleation of Ligand-Receptor Domains in Membrane Adhesion. Phys. Rev. Lett. 109:258101. 45. Farago, O., 2010. Fluctuation-induced attraction between adhesion sites of supported membranes. Phys. Rev. E 81:050902. 46. Pécréaux, J., H.-G. Döbereiner, J. Prost, J.-F. Joanny, and P. Bassereau, 2004. Refined contour analysis of giant unilamellar vesicles. Eur. Phys. J. E 13:277 -- 290. 47. Turlier, H., D. A. Fedosov, B. Audoly, T. Auth, N. S. Gov, C. Sykes, J.-F. Joanny, G. Gompper, and T. Betz, 2016. Equilibrium physics breakdown reveals the active nature of red blood cell flickering. Nat. Phys. . 16 Manuscript submitted to Biophysical Journal Static properties of a pinned membrane 48. Reister-Gottfried, E., K. Sengupta, B. Lorz, E. Sackmann, U. Seifert, and A.-S. Smith, 2008. Dynamics of Specific Vesicle-Substrate Adhesion: From Local Events to Global Dynamics. Phys. Rev. Lett. 101:208103:1 -- 4. 49. Deserno, M., 2015. Fluid lipid membranes: From differential geometry to curvature stresses. Chemistry and Physics of Lipids 185:11 -- 45. http://www.sciencedirect.com/science/article/pii/S000930841400053X, membrane mechanochemistry: From the molecular to the cellular scale. 50. Fenz, S. F., A.-S. Smith, R. Merkel, and K. Sengupta, 2011. Inter-membrane adhesion mediated by mobile linkers: Effect of receptor shortage. Soft Matter 7:952 -- 962. 51. Helfrich, W., 1978. Steric Interaction of Fluid Membranes in Multilayer Systems. Z. Naturforsch., A: Phys. Sci. 33:305. 52. Schmidt, D., C. Monzel, T. Bihr, R. Merkel, U. Seifert, K. Sengupta, and A.-S. Smith, 2014. Signature of a Nonharmonic Potential as Revealed from a Consistent Shape and Fluctuation Analysis of an Adherent Membrane. Phys. Rev. X 4:021023. 53. Benhamou, M., 2011. Primitive interactions between inclusions on a fluid membrane: The role of thermal fluctuations. The European Physical Journal E 34:79. https://doi.org/10.1140/epje/i2011-11079-6. 54. Helfrich, W., and R.-M. Servuss, 1984. Undulations, Steric Interaction and Cohesion of Fluid Membranes. Nuovo Cimento D 3:137 -- 151. 55. Lipowsky, R., 1991. The conformation of membranes. Nature 349:475 -- 481. 56. Lipowsky, R., 1995. Generic Interactions of Flexible Membranes. In R. Lipowsky, and E. Sackmann, editors, Structure and dynamics of membranes, Elsevier, chapter 11, 521 -- 602. 57. Rädler, J. O., T. J. Feder, H. H. Strey, and E. Sackmann, 1995. Fluctuation analysis of tension-controlled undulation forces between giant vesicles and solid substrates. Phys. Rev. E 51:4526 -- 4536. 58. Costa, J. A., and C. Brebbia, 1985. The boundary element method applied to plates on elastic foundations. Engineering Analysis 2:174 -- 183. http://www.sciencedirect.com/science/article/pii/0264682X85900292. 59. Chen, W., 2003. Boundary knot method for Laplace and biharmonic problems. CoRR cs.CE/0307061. http://arxiv. org/abs/cs.CE/0307061. 60. Bruinsma, R., A. Behrisch, and E. Sackmann, 2000. Adhesive switching of membranes: Experiment and theory. Phys. Rev. E 61:4253 -- 4267. 61. Limozin, L., and K. Sengupta, 2009. Quantitative Reflection Interference Contrast Microscopy (RICM) in Soft Matter and Cell Adhesion. ChemPhysChem 10:2752 -- 2768. 62. Goulian, M., R. Bruinsma, and P. Pincus, 1993. Long-Range Forces in Heterogeneous Fluid Membranes. EPL (Europhysics Letters) 23:155. http://stacks.iop.org/0295-5075/23/i=2/a=014. 63. Park, J.-M., and T. C. Lubensky, 1996. Interactions between membrane Inclusions on Fluctuating Membranes. Journal de Physique I 6:1217 -- 1235. http://www.edpsciences.org/10.1051/jp1:1996125. 64. Dommersnes, P. G., and J.-B. Fournier, 1999. N-body Study of Anisotropic Membrane Inclusions: Membrane Mediated Interactions and Ordered Aggregation. Eur. Phys. J. B 12:9 -- 12. 65. Weikl, T. R., M. M. Kozlov, and W. Helfrich, 1998. Interaction of conical membrane inclusions: Effect of lateral tension. Phys. Rev. E 57:6988 -- 6995. 66. Lin, H. K., R. Zandi, U. Mohideen, and L. P. Pryadko, 2011. Fluctuation-induced forces between inclusions in a fluid membrane under tension. Phys. Rev. Lett. 107:2 -- 6. 67. Bitbol, A. F., P. G. Dommersnes, and J. B. Fournier, 2010. Fluctuations of the Casimir-like force between two membrane inclusions. Phys. Rev. E 81:29 -- 32. 68. Müller, M. M., M. Deserno, and J. Guven, 2005. Interface-mediated interactions between particles: A geometrical approach. Phys. Rev. E 72:061407. https://link.aps.org/doi/10.1103/PhysRevE.72.061407. Manuscript submitted to Biophysical Journal 17 Janeš et. al 69. Sigurdsson, J. K., F. L. H. Brown, and P. J. Atzberger, 2013. Hybrid continuum-particle method for fluctuating lipid bilayer membranes with diffusing protein inclusions. Journal of Computational Physics 252:65 -- 85. http: //www.sciencedirect.com/science/article/pii/S0021999113004403. 70. Abramowitz, M., and I. Stegun, editors, 1965. Handbook of Mathematical Functions: with Formulas, Graphs, and Mathematical Tables. Dover Publications. 71. Baddour, N., 2009. Operational and convolution properties of two-dimensional Fourier transforms in polar coordinates. J. Opt. Soc. Am. A 26:1767 -- 1777. http://josaa.osa.org/abstract.cfm?URI=josaa-26-8-1767. 72. Olver, F. W., D. W. Lozier, R. F. Boisvert, and C. W. Clark, 2010. NIST Handbook of Mathematical Functions. Cambridge University Press, New York, NY, USA, 1st edition. 18 Manuscript submitted to Biophysical Journal
1111.2143
1
1111
2011-11-09T09:07:52
Challenging packaging limits and infectivity of phage {\lambda}
[ "physics.bio-ph", "q-bio.SC" ]
The terminase motors of bacteriophages have been shown to be among the strongest active machines in the biomolecular world, being able to package several tens of kilobase pairs of viral genome into a capsid within minutes. Yet these motors are hindered at the end of the packaging process by the progressive build-up of a force resisting packaging associated with already packaged DNA. In this experimental work, we raise the issue of what sets the upper limit on the length of the genome that can be packaged by the terminase motor of phage {\lambda} and still yield infectious virions, and the conditions under which this can be efficiently performed. Using a packaging strategy developed in our laboratory of building phage {\lambda} from scratch, together with plaque assay monitoring, we have been able to show that the terminase motor of phage {\lambda} is able to produce infectious particles with up to 110% of the wild-type (WT) {\lambda}-DNA length. However, the phage production rate, and thus the infectivity, decreased exponentially with increasing DNA length, and was a factor of 103 lower for the 110% {\lambda}-DNA phage. Interestingly, our in vitro strategy was still efficient in fully packaging phages with DNA lengths as high as 114% of the WT length, but these viruses were unable to infect bacterial cells efficiently. Further, we demonstrated that the phage production rate is modulated by the presence of multivalent ionic species. The biological consequences of these finding are discussed.
physics.bio-ph
physics
Challenging packaging limits and infectivity of phage λ Elmar Nurmemmedov1,a, Martin Castelnovo2, Elizabeth Medina3, Carlos Enrique Catalano3, and Alex Evilevitch1,4* 1) Department of Biochemistry, Center for Chemistry and Chemical Engineering, Lund University, Box 124, SE-221 00, Lund, Sweden 2) Laboratoire de Physique, Ecole Normale Superieure de Lyon, 46 Allée d'Italie, 69364 Lyon Cedex 07, France 3) Department of Medicinal Chemistry, University of Washington School of Pharmacy, H172 Health Sciences Building, Box 357610, Seattle, WA 98195, USA 4) Department of Physics, Carnegie Mellon University, 5000 Forbes Ave, Pittsburgh, PA 15213, USA * To whom correspondence should be addressed e-mail: [email protected] a Present address: Division of Hematology/Oncology, Children’s Hospital Boston, Harvard Medical School, Boston, MA 02115, USA. ABSTRACT The terminase motors of bacteriophages have been shown to be among the strongest active machines in the biomolecular world, being able to package several tens of kilobase pairs of viral genome into a capsid within minutes. Yet these motors are hindered at the end of the packaging process by the progressive build-up of a force resisting packaging associated with already packaged DNA. In this experimental work, we raise the issue of what sets the upper limit on the length of the genome that can be packaged by the terminase motor of phage λ and still yield infectious virions, and the conditions under which this can be efficiently performed. Using a packaging strategy developed in our laboratory of building phage λ from scratch, together with plaque assay monitoring, we have been able to show that the terminase motor of phage λ is able to produce infectious particles with up to 110% of the wild-type (WT) λ-DNA length. However, the phage production rate, and thus the infectivity, decreased exponentially with increasing DNA length, and was a factor of 103 lower for the 110% λ-DNA phage. Interestingly, our in vitro strategy was still efficient in fully packaging phages with DNA lengths as high as 114% of the WT length, but these viruses were unable to infect bacterial cells efficiently. Further, we demonstrated that the phage production rate is modulated by the presence of multivalent ionic 1 species. The biological consequences of these finding are discussed. Keywords: bacteriophage lambda, pressure, DNA packaging, terminase INTRODUCTION Viruses are among the simplest biological organisms. They typically consist of a viral genome within a protein container, called a capsid, whose function is to protect the genome and to provide a specific strategy for the early steps of infection. Despite their simplicity, viruses exhibit extreme diversity and are highly prone to mutations. This limits the efficiency of current anti-viral therapies, which are often focused on the specificity of viral replication. The past decade has seen the emergence of a new interdisciplinary approach, called “Physical Virology”, providing tremendous opportunities for the elucidation of the general physical mechanisms involved in virus development and infection. For example, in double-stranded (ds) DNA viruses, the genome is often hundreds of times longer than the dimensions of the capsid into which it is packaged, and this results in high internal pressure due to the repulsive negative charges on the densely packaged DNA. Indeed, high mechanical pressure in dsDNA bacterial viruses (phages) has recently been demonstrated (1, 2). During infection, the phage binds to the receptor on the bacterial cell surface and the pressure injects the genome into the cell, at least in part, while the empty capsid remains outside the cell. This initiates a series of events, eventually leading to hundreds of viral genomes, pre-assembled capsids and viral tails that are the direct precursors to the assembled, infectious virus. During phage assembly, a motor complex (the terminase enzyme), specifically recognizes viral DNA then binds to the portal complex situated at a unique vertex of the icosahedral capsid; the portal is a ring-like structure that provides a hole through which DNA enters the capsid during assembly and exits during infection (3-6). The terminase motors, which utilize ATP hydrolysis to drive DNA into the capsid, are among the most powerful biological motors characterized to date (3-6). For instance, the bacteriophage λ terminase motor packages DNA to near-crystalline density, which generates an internal pressure in excess of 20 atmospheres. Commensurate with this requirement, the λ terminase motor exerts forces greater than 50 piconewtons during the terminal stages of genome packaging (2, 4, 7-9). Similar features are observed in many bacteriophages and the eukaryotic herpes virus groups. In the latter case, genome packaging occurs in the nucleus of an infected cell, following a pathway that is remarkably similar to that of the phage system, i.e., a terminase motor specifically 2 packages viral DNA into the interior of a pre-assembled procapsid through a unique portal vertex (10-12). Using phage λ, Evilevitch and co-workers provided the first experimental in vitro evidence that internal pressure, often greater than 50 atm, ejects viral DNA from the capsid into the host cell (1, 13). Since the internal capsid pressure is mainly the result of electrostatic repulsion between neighboring DNA strands, we found that the pressure is markedly reduced in the presence of multivalent cations (due to the screening of negative charges on the DNA), since capsids are permeable to water and salts (14). Similar results have been observed in single- molecule laser tweezer experiments (4). Consistent with this, we also found that internal pressure increases strongly with increasing length of the DNA packaged into the capsid, as the negative charge density increases (13). This may explain the observation in vivo that a λ genome can only be 8.8 % longer than the wild-type λ genome and still yield an infectious virus in vivo (15). DNA packaging is used for molecular cloning in so-called cell extract “packaging kits”, where non-viral genes are inserted into the phage chromosome, packaged in vitro, delivered and expressed in bacterial cells (16). This technique has played a revolutionary role in the development of molecular biology and nanotechnology (17). However, while powerful, the size of an inserted gene is limited by the force of the terminase motor and its capacity to package oversized genomes into the capsid interior. During the early steps of the packaging process, the repulsive force exerted by the packaged DNA is smaller than the packaging force generated by the terminase motor. However, as the length of packaged DNA increases, repulsive forces increase and eventually exceed the maximum force that the motor can generate. Thus, the efficiency of viral packaging and virion synthesis are strongly dependent on the DNA length, ionic strength conditions, and types of cations included in the reaction mixture. Mg2+ and spermidine3+ can stabilize condensed DNA by reducing DNA–DNA repulsion. This may influence the rate of packaging and thus the number of successful packaging events. To our knowledge, no systematic or quantitative in vitro measurements have been made to explore the dependence of viral infectivity on the length of the packaged genome and the salt concentration, and the consequences for both controlled laboratory viral assembly (e.g. genetic engineering) and in vivo virus assembly. Indeed, this is not possible in the commercially available packaging kits that utilize crude cell extracts. Catalano and co-workers recently developed an in vitro virus assembly system in which 3 an infectious virus was assembled from purified components in a rigorously defined biochemical reaction mixture (5). This allows detailed and quantitative interrogation of each step of the assembly pathway and manipulation of the reaction conditions in a strictly controlled manner. In this work, we exploit this system to investigate the correlation between the internal genome pressure, governed by both the genome length and ionic conditions, and virus infectivity. Our goal was to determine the maximum length of the packaged DNA in phage λ by packaging oversized DNA, relative to the wild-type (WT) λ genome length of 48,502 bp, under various ionic strength conditions. The length of DNA was systematically varied by incorporation of non- specific plasmid DNA into the WT λ genome, and the packaging rate and efficiency were determined by several in vitro and in vivo assays. The results of these studies are discussed with respect to viral genome packaging in vivo and molecular cloning in vitro. RESULTS We constructed three oversized phage λ chromosomes: 51,206 bp (105%), 53,892 bp (110%) and 55,340 bp (114%) long, relative to the WT λ-genome length of 48,520 bp (corresponding to 100%), as described in Materials and Methods section. Briefly, we inserted linearized bacterial plasmids of defined lengths into the XbaI-digested λ chromosome (XbaI is a unique restriction site located at bp 24,511 of the λ sequence), without disturbing essential λ genes (Figure 1). The inserts contain two markers that allow detection of viral DNA entry into the host cell; lacZ for blue-clear bacterial colony detection, and ampR for ampicillin resistance (Figure 2). The oversized λ genomes were used as packaging substrates in our defined in vitro virus assembly system, as described previously (5). The packaging buffer contained 20 mM Tris- HCl, pH 7.5, 4.5 mM MgCl2 and 2 mM spermidine3+ (buffer 0), which is optimal for the assembly of an infectious virus using a WT (100%) λ genome (5). As anticipated, infection of E. coli JM83 cells (ampR-, lacZ-) with the in vitro assembled WT virus did not afford any bacterial colonies when plated on ampicillin-containing plates, because the WT genome does not contain the ampR gene. We next examined particle assembly and the infectivity of viruses assembled with each of the oversized genomes. As before, the genomes were used as packaging substrates in our in vitro virus assembly system, and E. coli JM83 was infected with the in vitro assembled virus. Successful packaging of the oversized genome into the procapsid, addition of a viral tail, and subsequent injection of the viral DNA into the host cell by the infectious assembled virus 4 was indicated by the presence of blue colonies (lacZ) that grow in the presence of ampicillin (ampR) (Figure 2). The data presented in Table 1 demonstrate that oversized genomes up to 114% of the WT genome length can be successfully packaged in vitro, and that the virus is able to inject the oversized DNA into the host cell to render them ampicillin-resistant. Importantly, control experiments, in which ATP was omitted from the virus assembly reaction, and thus cannot package DNA, did not yield any colonies on ampicillin-containing plates, indicating that virus particle assembly is required for DNA entry into the cell. Finally, we investigated whether salt concentration affected in vitro virus assembly or infectivity, as follows. Each of the genomes was used as packaging substrate, as described above, except that spermidine was omitted from the reaction (buffer 1), or the concentration of Mg2+ was reduced to 2 mM (buffer 2). (The terminase packaging motor has a strict divalent metal requirement, which precluded the complete omission of Mg2+ from the reaction mixture (5)). All the oversized genomes could be packaged into infectious viruses in vitro under all buffer conditions examined (see below). While the above results demonstrate that oversized genomes can be assembled into infectious viruses in vitro, they do not address the question of whether the oversized genomes can be replicated and packaged into infectious viruses in vivo. To directly address this question, we turned to a standard viral plaque assay. Each of the genomes was used to assemble an infectious virus, as described above. The in vitro assembled virus was then used to infect E. coli LE392 cells, and the number of infectious virus particles quantified by plating on a bacterial lawn (18). Each of the observed plaques results from the infection of a single cell with a virus particle (18), and provides a direct method of quantifying the number of infectious virus particles in the assembly mixture. Further, the injected genome must be competent for multiple rounds of replication and infectious virus particle assembly within the host cell for a plaque to be visible. Thus, the generation of a viral plaque requires that the genome is competent for the assembly of an infectious virus both in vitro and in vivo. The data presented in Table 1 demonstrate that the 105% and 110% oversized genomes provide a functional substrate for DNA replication and virus development in vivo, as evidenced by the appearance of viral plaques on the bacterial lawn. In contrast, the 114% genome substrate failed to yield visible plaques, despite the fact that the elongated genome can be packaged and subsequently delivered to the cell in vitro (Table 1). This indicates that a 114% long λ genome can be packaged under defined reaction conditions in vitro, but that the cellular cytoplasm does 5 not provide an environment conducive to the assembly of an infectious virus with this genome, at least over the multiple infection cycles required to visualize a plaque. We have thus identified the limit of λ genome length that can be efficiently packaged in vivo. While the number of plaques reflects the efficiency of virus assembly in vitro, the size of the viral plaque is related to the rate of virus particle assembly in vivo – the faster (more efficiently) the virus replicates in vivo, the larger the plaque. Figure 3 shows a direct comparison between the 100% and 110% λ genome phages plated on the same LE392 plate, which clearly demonstrates a significant difference in plaque size. As all other factors remain unchanged, and since genome packaging is the rate-limiting step in the concerted viral assembly process (19), we interpret the dramatic reduction in viral replication rate as a reflection of a decrease in the rate of genome packaging in vivo. This observation is however not trivial because there are more factors contributing to the plaque size at a given time. The above interpretation assumes that the input genome was faithfully replicated and packaged in vivo to yield infectious, over-packaged virions. We considered the possibility that the oversized genomes were successfully packaged in vitro and injected into the host cell, but the duplex length was somehow shortened during replication in vivo to generate a viable virus. To directly examine this possibility, the DNA was extracted from the plaques derived from packaging the 105% and 110% oversized genomes. The size of the virion DNA was confirmed with pulsed-field gel electrophoresis, and Figure 4 shows that the input genome is faithfully replicated and packaged into infectious virus particles in vivo. We next quantified the efficiency of virus assembly in vitro as a function of genome length. For these studies we utilized our standard reaction buffer 0 (20 mM Tris-HCl pH 7.5, 2 mM spermidine and 4.5 mM MgCl2), and the assembly reaction was allowed to proceed for 2 hours to ensure that it had reached completion. Virus assembly was quantified by plaque assay. The plates were incubated for 12 hours at 37°C, sufficient for the plaques to appear. As can be seen in Figure 5 (red dashed line), the number of plaques, and thus the efficiency of packaging, shows an exponential decrease with increasing packaged λ-DNA length. The number of assembled virions decreased exponentially by a factor of 100 for 5% extra packaged DNA length, relative to the WT. The titer fell by 1000 times when the λ-DNA length was increased from 100% to 110%, and would presumably be 10,000 times lower for the 114% λ-DNA sample. The latter would reduce the transfection efficiency to less than that required for detection by the 6 plaque assay. Similar relative drop in the packaging efficiency was also found for in vivo packaging in ref. (15), when increasing λ-DNA length from 105% to 108.8%. However, the initial drop in the efficiency when going from 100% to 105% λ-DNA length was smaller than in the present study. This is explained by the fact that salt conditions cannot be controlled in an in vivo packaging process unlike in our in vitro packaging study. As we have shown here, packaging efficiency will have a strong ionic dependence for smaller variation in the packaged DNA length around 100% WT λ-DNA length (see also discussion below). Finally, we quantified the effect of salt concentration on the efficiency of virus assembly in vitro. Specifically, we decreased the concentrations of spermidine3+ (Figure 5, green dashed line) and Mg2+ (Figure 5, blue dashed line) independently, to ensure that any observed effect was the result of a counterion-induced decrease in DNA–DNA repulsion, rather than the effect of a specific salt on the viral protein assembly during packaging. In both cases, decreasing the counterion concentration significantly decreased the yield of infectious viruses in vitro, with spermidine exhibiting slightly stronger effects (Figure 5). DISCUSSION We have examined the assembly of an infectious λ virus in vitro under rigorously defined conditions, and have demonstrated that both the rate and extent of particle assembly are strongly affected by the length of the genome substrate. We have further demonstrated that while very large DNA duplexes (114% of WT genome length) can be packaged into a particle and efficiently delivered to a host cell in vitro, there is a more stringent limit on virus development in vivo. What limits the efficiency of packaging over-sized genomes? Effect of DNA length on the packaging efficiency and viral infectivity. There is an exponential decrease in the yield of infectious viral particles assembled in vitro as a function of genome length. The only variable in the assembly reaction is the length of the DNA used as a packaging substrate, and we suggest two reasonable explanations of this finding. First, it might reflect differences in the rate of assembly of the packaging motor complex. The reactions are initiated by terminase, which must bind to the DNA, and the binary complex must then bind to the procapsid to initiate packaging. It is unlikely that a change in the genome length would affect the activation of the packaging motor once assembled. We consider, however, that the rate- limiting step in packaging initiation is related to the free diffusion of the DNA coil in the 7 solution in order to first find a terminase and a free procapsid. From classical polymer physics it can be estimated that the change in diffusion coefficient for two different DNA lengths, LWT and D(L) ≈ D(LWT ) LWT / L ν = 1 / 2 or 3 / 5 )ν , where the exponent of polymer statistics is L, scales as ( for an ideal Gaussian chain or a swollen chain (20). As a consequence, we anticipate a reduction of only 5% in the diffusion coefficient of the DNA coil when its length is increased from the WT € € length to 110% of the WT length. This modest change in diffusion coefficient is unlikely to be responsible for the large decrease in virus assembly (of the order of 103) observed in our experiments. An alternative explanation is that the observed decrease in the rate and extent of infectious particle assembly reflect a significant increase in the time required to package the oversized duplex, and a decrease in the efficiency of particle completion in the face of an over- packaged capsid, respectively (Figure 1). As we argue below, this exponential decay is related to the dynamic properties of the terminase packaging motor under an external load (i.e. the external load does not arise from the motor itself). The mechanical properties of viral packaging motors have been characterized experimentally at the single molecule level during the past decade (2). Using laser tweezer techniques, it has been shown for phage λ (4) and for phage φ29 (21), that if an external load challenges the active packaging process of the terminase motor, the leading order relation Fδ % ( v ≈ v 0 exp − v 0 is the * , where ' kT ) & δ is the step size of the motor, k is measured velocity in the absence of the load, and Boltzmann’s constant and T is the temperature. Note that this relation has been established € € experimentally as a dynamic intrinsic property of the motor, regardless of DNA filling inside the € capsid. The origin of this exponential behavior is intimately related to the decrease in power stroke efficiency of the terminase motor under an opposing external load (22) and the increased frequency of slips of the DNA as it is being packaged at higher loading forces (2). The equations require an irreversible step, which is the case here because cos-cleavage of λ-DNA is irreversible. In the absence of an external load, it has also been observed that packaging rate slows down dramatically after the first 30% of the WT λ-genome has been packaged. This is the signature of the internal force build-up that resists active packaging by the motor, arising from between the packaging rate v and this external load F scales as 8 T (L) ≅ T LWT ( ) + T (L) = € package DNA of length L, the DNA confinement in the capsid. Slowing of the packaging rate is qualitatively described by a Fint (L)δ % ( v (L) ≈ v 0 exp − * , where L is the packaged similar relation between velocity and force ' kT & ) Fint (L) is the internal force associated with the confinement of DNA of length L length and within the capsid. Therefore, we used this relation to estimate the average time, T(L), required to € exp Fint ( x )δ $ ' L∫ dx & ) kT 1 ( % . v 0 For DNA around the same length as that of WT (48.5kb) or larger, Fuller et al. (4) showed that the internal force is large (25 pN at 90% of WT DNA length). Assuming, for the € sake of simplicity, that the relation between the internal force and the genome length is linear above the WT DNA length in the range addressed in our experiments, the average time required to package a length L, that is larger than the WT length LWT, scales roughly as suggested both by the experimental results of Fuller et al. (4) and by DNA packaging models, such as the inverse spool model (23-26): kT exp A(L − LWT )δ % ( ' * kT & ) Av 0δ (Eq. 1) where A is the coefficient of the linear relationship between the internal force and the genome length. The dynamic properties of the terminase motor therefore predict an exponential increase in the time required to fully package oversized genomes. As a consequence, the rate of virus J (L) = 1 /T (L) will decay exponentially with the genome length. This conclusion assembly implicitly assumes that the rate-limiting step for the production of N viruses is the assembly rate, N ( L ) ≈ J ( L )Texp , where Texp is the duration of the packaging experiment. Provided that such that € the incubation time for in vitro phage packaging and assembly was 2 hours for all samples, and that the assembly components and enzymes retain their activity during this time period, € € exponential decay in the production rate leads to the exponential decay of the number of assembled, infectious phage particles, in agreement with the experimental results presented here. Infinitely long packaging times may not even be feasible in vivo as cells may run out of ATP required for packaging, and/or the packaging enzymes and viral DNA may be degraded. Indeed, the decrease in plaque size as a function of genome size indicates that the rate of particle assembly is also dramatically slower in vivo. Thus, a change in the packaged DNA length of 9 € even a few percent, accompanied by an increase in the internal pressure, will have dramatic consequences for viral reproduction and the spread of infection, decreasing by several orders of magnitude. Effect of salt on DNA packaging. The data presented here provide further information on the influence of ionic species on the efficiency of virus assembly under rigorously defined conditions. Our results are consistent with recent single-molecule laser tweezer studies showing that polyamines and Mg2+ decrease the internal force resisting packaging (7). Moreover, the single-molecule packaging rate is systematically faster in the presence of multivalent ions, at least when packaging a unit-length WT genome. The consequence of this observation is that the time required to package DNA up to WT genome length should be shorter in the presence of multivalent ions. Our data indicate that the time required to package DNA beyond its WT length, up to 110% of the natural genome, depends strongly on a combination of the dynamic properties of the motor and the force resisting packaging. Examination of Fuller’s measurements (7) suggests that the relation between the internal force and packaged DNA length around the WT length is still linear and is only slightly dependent on the ionic conditions. Therefore, the previous scaling estimation of the time required for packaging of oversized λ-DNA can be reproduced. These two observations lead to the following conclusion: the initial packaging of DNA up to WT genome length is faster in the presence of multivalent ions, but the duration of the later stages of packaging beyond the WT genome length should still increase exponentially with genome length. As a consequence, our approach predicts that the addition of multivalent ions will increase the phage production rate at constant genome length, while maintaining approximately the same exponential decay with increasing genome length. This matches our experimental data qualitatively (Figure 5) and further confirms our interpretation of the overpackaging data. Thus, the phage production rate is increased with more efficient cation induced reduction of DNA-DNA repulsive force resisting packaging. This is achieved by higher cation charge (as can be seen by comparing spermidine3+ and Mg2+, both naturally present in the E. coli cytoplasm) or an increase in the overall ionic strength (Figure 5). Ionic strength values, I, for all three buffers were 13.5 mM, 18 mM and 25.5 mM, respectively, excluding the contribution from the Tris buffer. It is interesting to note that with increasing DNA length, the effect of salt on the 10 packaging efficiency becomes weaker. It can be seen in Figure 5 that the addition of 2 mM spermidine3+ (with 4.5 mM Mg2+ in both buffers 0 and 1) increases the titer by ≈20 times for 100% λ-DNA length, but only by ≈8 times for 110% λ-DNA packaging. In Figure 5, it can be seen that increasing the ionic strength from 13.5 mM (buffer 1) to 18 mM (buffer 2), increases the titer by ≈1.5 times for 100% λ-DNA but there is no titer change for 110% λ-DNA. Thus, the phage titer curves at different ionic conditions shown in Figure 5 are converging with increasing DNA length.1 This observation is consistent with well-documented observations using osmotic stress experiments (27-29), where it was shown that DNA-DNA interactions in dense hexagonal phases of DNA (similar to DNA densities in overpackaged phage λ) are dominated by the hydration forces and have little dependency on the identity of the salt. Furthermore, high DNA packaging densities lead to several times higher counter-ion concentration at the interior of the capsid, compared to the bulk, making the internal forces less sensitive to the external ion content (30, 31). 1 It should be noted that y-axis, showing pfu/mL, is plotted on a log-scale, making the lines look essentially parallel. However, the differences in phage titer described above between different buffers at various DNA lengths are significant as all 3 exponential curves are asymptotically approaching zero. 11 CONCLUSIONS Elegant studies by Feiss and co-workers (15) have demonstrated that genomes as long as 1.08 could be packaged from di-lysogens in vivo, which sets an upper limit on productive virus development in the cell. The goal of this work was to (i) define the physical limit of DNA that can be inserted into a capsid by the terminase motor and (ii) to elucidate the role of the internal genome pressure on virion synthesis and viral infectivity. These studies take advantage of biochemically defined in vitro virus assembly assay recently developed in the Catalano lab. The genome pressure was varied by changing two parameters: DNA length and salt condition in the virus packaging-assembly buffer. The key to these measurements was our recently developed system for the in vitro assembly of infectious phage λ, using individually purified components in a rigorously defined biochemical assay (5). For the first time, we were able to quantitatively measure a dramatic reduction in the phage production rate associated with the packaging of DNA longer than the WT genome in the phage λ capsid. These measurements reveal important biological implications. The first is that it is possible to package DNA up to 114% of the WT genome length for phage λ. Moreover, by using different ionic conditions, we demonstrated that overpackaging is a rather robust feature. The results of our investigations have implications with respect to the presence and the origin of a maximum length of DNA that can be packaged within a viral capsid. Indeed, we found that longer DNA than WT genome length can be packaged. However, the rate of packaging, and thus the number of phages produced in a given time, decreases exponentially with increasing genome length. More precisely, a 1% increase in the length of the packaged genome above WT, leads to ten-fold decrease in the viral titer. Moreover, since salt decreases the internal pressure, adding extra salt increases the titer. With higher packaging density (above 110% of WT λ-DNA) the hydration force dominates the DNA-DNA interaction and the pressure is only weakly affected by the addition of counterions, making the packaging rate essentially insensitive to salt conditions. As a consequence, we found that the maximum length of DNA that could be packaged is not limited by the salt conditions of the virus assembly environment. Instead, it is limited by the maximum force that the terminase motor can exert. Remarkably, we also found that maximum length of DNA that can be packaged in vivo is between 110 and 114% of WT λ-DNA, however the packaging efficiency was very low. These observations illustrate that the dynamics of the viral packaging motor is 12 exponentially dependent on the genome length. The motor packaging rate approaches zero asymptotically as the internal pressure increases. Thus, there is no mechanical limit on the length of DNA that can be packaged (assuming that the capsid can withstand any force). This has interesting implications on DNA packaging by the “headful” phages. These viruses initiate packaging at a specific “pac” site in the concatemer, but package longer than unit length duplexes (102-104% genome length). The mechanism that defines the downstream cut site remains unclear, but slowing of the terminase packaging motor certainly plays a major role. Indeed, a simple model is that upon packaging a headful of DNA, the terminase motor slows to essentially zero as the pressure of the packaged DNA overcomes the power of the motor. This provides time for the slow endonuclease activity of terminase to cut the duplex to complete the packaging process. In the case of unit-length genome packaging viruses such as λ, the packaging time becomes infinitely long as DNA length exceeds WT, preventing viruses from assembling within reasonable time (or at all) to complete their replication cycle. This suggests that the length of the packaged genome has been evolutionarily optimized to packaging densities where the effect of salt on the motor efficiency is minimized, making packaging more robust and less prone to changes in the cellular environment. Most importantly, these findings demonstrate that slight changes in the internal pressure induced by variation in the DNA length, will have fatal consequences for virus synthesis and the spread of infectious particles. This was demonstrated by the dramatic reduction in the number (efficiency) and size (rate) of plaques. These observations strengthen the role of internal pressure in the infectivity process, in agreement with several in vivo studies (32, 33). This study also suggests new strategies for interfering with viral infectivity through small changes in the internal genome pressure in the case of motor-packaged DNA viruses. At the same time, the fact that overpackaging is possible in various environments might be considered as a potential evolutionary advantage: as phage λ would therefore have the ability to modulate its genome length, since both its terminase motor characteristics and its host (the bacteria) are capable of packaging a genome larger than the WT length. More precisely, our finding of an effective upper limit for packaging of the oversized 114% WT DNA means that modulation of genome length is possible up to an extra length of 6.8 kb. Considering that the typical average size of a gene is of the order of 1 kb, phage λ therefore has the ability to increase 13 its genome by at least one or two genes. This is a remarkable feature with respect to the adaptation of the virus to its environment. Based on recent results obtained in our studies (34, 35) and those at other labs (4), on the physical properties of phage λ, we speculate that λ WT length is the optimal result of a balance between a large number of genes, in order to function properly, a long genome length in order to be able to efficiently initiate passive ejection of the genome into the host bacteria, and a genome length short enough that efficient virus production is compatible with the time frame of the cell cycle. MATERIALS AND METHODS Construction of oversized phage λ chromosomes. Oversized phage λ chromosomes, 105%, 110% and 114% of the length of the wild type, were constructed by insertion of linearized bacterial plasmids of defined length into the XbaI-digested λ chromosome. XbaI is a unique restriction site that is located at 24511 base pairs in an inert gene that seemingly does not affect phage replication. 200 µg of λ DNA (dam- dcm-) was digested with 20 units XbaI at 37 ºC during a period of 12 hours. Digested λ DNA was treated with calf intestinal alkaline phosphatase at 37 ºC for 5 hours. The λ DNA was then purified using a phenol-chloroform purification technique. Bacterial plasmid pUC19 (AmpR and LacZ+, 2686 base pairs) was used for construction of the 105% and 110% λ chromosomes. Firstly, it was made compatible for the ligation with the digested λ DNA. An adaptor was inserted into its unique PscI site. The adaptor (5'- CATGGCTAGCAACCTAGGAAACTAGT-3' and 5'-CATGACTAGTTTCCTAGGTT GCTAGC -3') carries three restriction sites (NheI, AvrII and SpeI), all of which produce sticky ends compatible with those of XbaI. Thus, modified-pUC19 was generated. This was consequently used for the construction of double-pUC19, which is basically two fused copies of the same plasmid. Modified-pUC19 was digested with NheI, purified from 0.8% agarose gel and ligated with itself. Ligation products, amplified in E. coli, were enzymatically linearized and separated on 0.8% agarose gel electrophoresis. The DNA fragment corresponding to twice the size of pUC19 was extracted, purified and circularized with T4 DNA ligase. Thus double-pUC19 (5372 base pairs) was created. For the construction of the 114% λ chromosome, plasmid pSV-b- galactosidase (AmpR and LacZ+, 6820 base pairs) was used. This plasmid has a unique XbaI site, which is conveniently located in an inert region. 14 In order to construct the 105%, 110% and 114% λ chromosomes, 50 µg of each of the modified-pUC19, double-pUC19 and pSV-β-galactosidase plasmids were digested separately with 10 units of SpeI (for modified-pUC19 and double-pUC19) and XbaI (for pSV-β- galactosidase) at 37 ºC over the course of 12 hours. DNA was subsequently separated on 0.8% agarose gel electrophoresis, and fragments corresponding to 2686, 5372 and 6820 base pairs were extracted and purified. Ligation reactions were subsequently performed with XbaI-digested and dephosphorylated λ DNA. 0.8 µg of λ DNA was used per 50 µl of ligation reaction, into which 1 unit of T4 DNA ligase was introduced. A molar ratio of 1:8 (1 for λ DNA, 8 for plasmid) was employed. Reactions were performed at 16 °C for 12 hours. Ligation was accompanied by four negative control reactions, in which the reaction components were sequentially removed: 1) no plasmid insert, 2) no T4 ligase, 3) no λ DNA, and 4) no λ DNA and no T4 ligase. The phenol-chloroform-purified total DNA product from ligation reactions was used as packaging substrate. Assembly of oversized λ–DNA phages. Phage λ was built from scratch from its recombinantly produced protein components in a two-step process. In comparison to the initially described method (1), we have slightly modified this process in terms of DNA and packaging buffer. In the packaging step, the following viral components were sequentially mixed in 25 µl reaction in a controlled ionic environment: 8.48 mM ATP, 95 µM gpD, 424 nM gpFI, 2 µM IHF, 208 nM procapsids, 1 µM terminase and DNA. The packaging process was performed under three different sets of ionic conditions: Buffer 0 20 mM Tris-HCl pH 7.5, 2 mM spermidine and 4.5 mM MgCl2 Buffer 1 Buffer 2 20 mM Tris-HCl pH 7.5 and 4.5 mM MgCl2 20 mM Tris-HCl pH 7.5, 2 mM spermidine and 2 mM MgCl2 The quantities of DNA used for this step were empirically determined (as described below). After incubation for 2 hours at 25 ºC, the packaging process was considered complete. Prior to the subsequent assembly step, the viral tail segment was constructed by mixing 417 µM gpFII, 413 µM gpW and 320 nM tail in 10 µl reaction, with 20 mM Tris-HCl pH 7.5, 2 mM spermidine, 3 mM MgCl2, 7 mM β-mercaptoethanol and 25 mM potassium glutamate. This mixture was then added to the packaged viral shells from the previous step and incubated for 2 hours at 37 ºC. This step resulted in a fully infective phage. Subsequently, the complete λ phage 15 was used for in vivo assays. Positive controls were similarly prepared using wild-type λ DNA. In vivo assays using E. coli In order to test our viral assembly system, we performed phenol-chloroform extraction of DNA from the packaged viral heads (directly after the packaging step). Extracted DNA was separated on 0.8% agarose gel for qualitative determination. However, this approach was successful only for the wild-type phage, as the DNA yield from the overpackaged phages was too low to be detected at this stage. Therefore, we performed two complimentary in vivo tests. Blue-clear screening was performed in order to confirm the integration of the plasmid DNA into the phage λ chromosome. Assembled phages were used to infect E. coli JM83 (AmpR- LacZ-), grown to an OD600 of 1.0 in LB medium supplemented with 10 mM MgSO4 and 0.2% (w/v) maltose. Infected cells were spread on LB agar plates containing 100 µg/ml ampicillin, 40 µg/ml X-gal and 0.5 mM IPTG (isopropyl β-D-1-thiogalactopyranoside). Plates were incubated at 37 ºC for 12 hours. E. coli JM83 infected with overpackaged phage λ yielded blue colonies, whereas E. coli JM83 infected with the wild-type phage λ did not yield any colonies. This in vivo test clearly demonstrated that the presence of phage-delivered AmpR and LacZ genes provides E. coli JM83 with ampicillin resistance and blue pigmentation. It is thus an indicator of the successful construction of oversized phage λ chromosomes. Plaque assay was performed for quantitative determination of packaging efficiency and in vivo infectivity of the overpackaged phages. Assembled phages were diluted 10, 100 and 500 times and mixed with 100 µl E. coli LE392, which was grown to an OD600 of 1.0 in LB medium supplemented with 10 mM MgSO4 and 0.2% (w/v) maltose. Infected cells were spread on LB agar plates. The plates were incubated at 37 ºC for 12 hours. Pulsed-field gel electrophoresis was used to visualize the DNA and assess the difference in size, using CHEF DR II electrophoresis equipment (Bio-Rad). Purified DNA samples were loaded into 1% agarose gel. The gel was run for 26.5 hours in 0.5x TBE buffer using the following conditions: initial switch of 2.9 seconds, final switch of 4.5 seconds and voltage of 6 V/cm. The gel was later stained with SYBR Gold nucleic acid dye and visualized under 302 nm UV. We observed a clear size difference between wild-type, 105% and 110% λ DNA bands. DNA saturation experiments: The amount of DNA used for the assembly of overpackaged λ phages affected the final yield, i.e. the number of plaques, as was previously shown (1). Therefore, we had to establish saturation levels for 105% and 110% DNA empirically. Since the 16 ligation product used as packaging substrate was a mixture of various DNA products, including the oversized DNA product, we used weight units (µg) of total DNA used for saturation level experiments. We designed phage assembly experiments (as described above) where various weight units, between 3 and 40 µg, of DNA substrate were used. Packaging buffer 0 was used for this experiment. These phages were subsequently used for plaque assay, as described above. Plaque numbers were counted and used to estimate the DNA load per packaging reaction that would ensure the oversaturation of packaging mixture with the DNA substrate. These saturation levels were later used in further experiments for the determination of the dependence of packaging efficiency on the ionic strength. ACKNOWLEDGMENTS: We would like to thank William Gelbart and Charles Knobler for fruitful discussions and inspiration, Lars-Olof Hedén for tremendous help with the microbiological part of this project, and John Pettersson for technical help. This work was supported by Carl Trygger’s Foundation grant to AE and the Swedish Research Council (VR) grant to AE (grant # 2008-726). REFERENCES 1. Evilevitch A, Lavelle L, Knobler CM, Raspaud E, & Gelbart WM (2003) Osmotic pressure inhibition of DNA ejection from phage. Proc. Nat. Acad. Sci. USA 100(16):9292–9295. Smith DE, et al. (2001) The bacteriophage phi 29 portal motor can package DNA against 2. a large internal force. Nature 413(6857):748-752. 3. Catalano CE, Cue D, & Feiss M (1995) Virus DNA packaging: the strategy used by phage lambda. (Translated from eng) Mol Microbiol 16(6):1075-1086 (in eng). Fuller DN, et al. (2007) Measurements of single DNA molecule packaging dynamics in 4. bacteriophage lambda reveal high forces, high motor processivity, and capsid transformations. (Translated from eng) J Mol Biol 373(5):1113-1122 (in eng). 5. Gaussier H, Yang Q, & Catalano CE (2006) Building a virus from scratch: assembly of an infectious virus using purified components in a rigorously defined biochemical assay system. J Mol Biol 357(4):1154-1166. 6. Rao VB & Feiss M (2008) The bacteriophage DNA packaging motor. (Translated from eng) Annu Rev Genet 42:647-681 (in eng). 17 Fuller DN, et al. (2007) Ionic effects on viral DNA packaging and portal motor function 7. in bacteriophage phi 29. Proc Natl Acad Sci U S A 104(27):11245-11250. 8. Fuller DN, Raymer DM, Kottadiel VI, Rao VB, & Smith DE (2007) Single phage T4 DNA packaging motors exhibit large force generation, high velocity, and dynamic variability. Proc Natl Acad Sci U S A 104(43):16868-16873. Rickgauer JP, et al. (2008) Portal motor velocity and internal force resisting viral DNA 9. packaging in bacteriophage phi29. Biophys J 94(1):159-167. Cardone G, et al. (2007) Visualization of the herpes simplex virus portal in situ by cryo- 10. electron tomography. Virology 361(2):426-434. 11. Newcomb WW, Booy FP, & Brown JC (2007) Uncoating the herpes simplex virus genome. J Mol Biol 370(4):633-642. Booy FP, et al. (1991) Liquid-crystalline, phage-like packing of encapsidated DNA in 12. herpes simplex virus. (Translated from eng) Cell 64(5):1007-1015 (in eng). Grayson P, et al. (2006) The effect of genome length on ejection forces in bacteriophage 13. lambda. Virology 348(2):430-436. Evilevitch A, et al. (2007) Effects of Salt Concentrations and Bending Energy on the 14. Extent of Ejection of Phage Genomes. Biophys J 15. Feiss M, Fisher RA, Crayton MA, & Egner C (1977) Packaging of the bacteriophage lambda chromosome: effect of chromosome length. Virology 77(1):281-293. 16. Hohn B & Murray K (1977) Packaging recombinant DNA molecules into bacteriophage particles in vitro. (Translated from eng) Proc Natl Acad Sci U S A 74(8):3259-3263 (in eng). Glick BR & Pasternak JJ (1998) Molecular Biotechnology (ASM Press) 2nd ed Ed. 17. T. Maniatis, E. F. Fritsch, J. Sambrook (1983) Molecular Cloning A Laboratory Manual 18. (Cold Spring Harbor Laboratory) seventh Ed. Hendrix RW ed (1983) Lambda II (Cold Spring Harbor, N.Y. : Cold Spring Harbor 19. Laboratory). Doi M & Edwards SF (1986) The theory of polymer dynamics (Clarendon Press, Oxford) 20. pp xiii, 391 p. Chemla YR, et al. (2005) Mechanism of force generation of a viral DNA packaging 21. motor. Cell 122(5):683-692. Jülicher F, Ajdari A, & Prost J (1997) Modeling molecular motors. Rev. Mod. Phys. 22. 18 69:1269-1282. 23. Kindt J, Tzlil S, Ben-Shaul A, & Gelbart WM (2001) DNA packaging and ejection forces in bacteriophage. Proceedings of the National Academy of Sciences of the United States of America 98(24):13671-13674. 24. Tzlil S, Kindt JT, Gelbart WM, & Ben-Shaul A (2003) Forces and pressures in DNA packaging and release from viral capsids. Biophys. J. 84:1616-1627. 25. Purohit PK, Kondev J, & Phillips R (2003) Mechanics of DNA packaging in viruses. Proc. Nat. Acad. Sci. USA 100(6):3173-3178. Purohit PK, et al. (2005) Forces during bacteriophage DNA packaging and ejection. 26. Biophys J 88(2):851-866. 27. Leikin S, Rau DC, & Parsegian VA (1991) Measured entropy and enthalpy of hydration as a function of distance between DNA double helices. Physical Review. A 44(8):5272-5278. Parsegian VA & Rau DC (1984) Water near intracellular surfaces. J Cell Biol 99(1 Pt 28. 2):196s-200s. 29. Rau DC & Parsegian VA (1992) Direct measurement of temperature-dependent solvation forces between DNA double helices. Biophys J 61(1):260-271. 30. Nevsten P, Evilevitch A, & Wallenberg R (2011) Chemical mapping of DNA and counter-ion content inside phage by energy-filtered TEM. J of Biol Phys, Online First DOI 10.1007/s10867-011-9234-8. 31. Li Z, Wu J, & Wang ZG (2008) Osmotic pressure and packaging structure of caged DNA. (Translated from eng) Biophys J 94(3):737-746 (in eng). 32. Koster S, Evilevitch A, Jeembaeva M, & Weitz DA (2009) Influence of internal capsid pressure on viral infection by phage lambda. (Translated from eng) Biophys J 97(6):1525-1529 (in eng). 33. De Paepe M & Taddei F (2006) Viruses' life history: towards a mechanistic basis of a trade-off between survival and reproduction among phages. (Translated from eng) PLoS Biol 4(7):e193 (in eng). 34. Ivanovska I, Wuite G, Jonsson B, & Evilevitch A (2007) Internal DNA pressure modifies stability of WT phage. Proc Natl Acad Sci U S A 104(23):9603-9608. Evilevitch A, et al. (2011) Effects of salts on internal DNA pressure and mechanical 35. properties of phage capsids. (Translated from eng) J Mol Biol 405(1):18-23 (in eng). 19 FIGURE LEGENDS: Figure 1: Over-sized phage λ chromosomes of length 105%, 110% and 114% were built by insertion of linearized bacterial plasmids of defined length into the XbaI-digested λ chromosome. All plasmids contained AmpR and LacZ genes providing ampicillin resistance and blue pigmentation upon infection of E.coli JM83. It is an indicator of successful construction and packaging of oversized phage λ chromosomes in vitro. In parallel, efficiency of the in vivo phage assembly after the bacterial cell infection is monitored by plaque assay, revealing the number of assembled and infectious phage particles determined by the number of plaques per plate. Figure 2: Blue-clear screening was performed in order to confirm integration of the plasmid DNA into the phage λ chromosome. Assembled phages were used to infect E.coli JM83 (AmpR- LacZ-). Infected cells were spread on LB-agar plates. E.coli JM83 infected with overpackaged phage λ yielded blue colonies, whereas E.coli JM83 infected with wild-type phage λ did not yield any colonies. This in vitro assembly test clearly demonstrated that presence of phage- delivered AmpR and LacZ genes provides ampicillin resistance and blue pigmentation to E.coli JM83 (Petri dish photo is shown next to the cartoon). Plaque assay was performed as a quantitative test of the virion assembly efficiency in vivo. Assembled phages were mixed with E.coli LE392. Infected cells were spread on LB-agar plates and after 12 hours incubation at 37 °C plaques could be observed (Petri dish photo is shown next to the cartoon). Figure 3: Plaque assay for WT (100%) and 110% λ-DNA phages on E. coli LE392 plate, illustrating large plaques for 100% DNA phage and small plaques for 110% DNA phage. Figure 4: Pulsed-field gel electrophoresis of DNA extracted from plaques carrying oversized λ genomes. Figure 5: Efficiency of packaging monitored by plaque assay. pfu/mL is plotted on a log-scale versus % of packaged DNA length (100% corresponds to the wild-type λ-DNA length). 20 in vitro virus assembly in vivo Procapsid Terminase motor λ genome WT = 100% = 48,502 bp 106% = 51,206 bp 111% = 53,892 bp 114% = 55,340 bp Infect E. coli ADP ATP gpW gpFII Tail Infectious VIrus Infection Assay ~ ampicillin resistant, blue bacterial colonies -> efficiency of virus assembly in vitro -> efficiency of DNA injection into host cell Plaque Assay ~ viral plaques on a bacterial lawn -> efficiency of virus replication in vivo (plaque number) -> rate of virus replication in vivo (plaque size) ! genome Infect E. coli Virus Particle Assembly Innoculate Bacteria Ampicillin Plates Virus Assembly Reaction Mix Blue Bacterial Colonies Indicate Particle Assembly and Host Cell Infection in Vitro Plate Bacterial Lawn (plaque assay) •Plaque Number (pfu) Reflects Assembly Efficiency in Vitro •Plaque Size Reflects Replication Rate in Vivo 107 106 105 104 1000 100 L m / u f p (I=25.5 mM) Buffer 0: 2 mM Sp3+, 4.5 mM MgCl 2 (I=18 mM) Buffer 2: 2 mM Sp3+, 2 mM MgCl 2 (I=13.5 mM) Buffer 1: 0 mM Sp3+, 4.5 mM MgCl 2 98 100 102 104 106 108 % packaged DNA 110 112
1007.1745
2
1007
2011-08-09T13:37:40
The mean velocity of two-state models of molecular motor
[ "physics.bio-ph", "math-ph", "math-ph" ]
The motion of molecular motor is essential to the biophysical functioning of living cells. In principle, this motion can be regraded as a multiple chemical states process. In which, the molecular motor can jump between different chemical states, and in each chemical state, the motor moves forward or backward in a corresponding potential. So, mathematically, the motion of molecular motor can be described by several coupled one-dimensional hopping models or by several coupled Fokker-Planck equations. To know the basic properties of molecular motor, in this paper, we will give detailed analysis about the simplest cases: in which there are only two chemical states. Actually, many of the existing models, such as the flashing ratchet model, can be regarded as a two-state model. From the explicit expression of the mean velocity, we find that the mean velocity of molecular motor might be nonzero even if the potential in each state is periodic, which means that there is no energy input to the molecular motor in each of the two states. At the same time, the mean velocity might be zero even if there is energy input to the molecular motor. Generally, the velocity of molecular motor depends not only on the potentials (or corresponding forward and backward transition rates) in the two states, but also on the transition rates between the two chemical states.
physics.bio-ph
physics
The mean velocity of two-state models of molecular motor Yunxin Zhang∗ Shanghai Key Laboratory for Contemporary Applied Mathematics, Centre for Computational System Biology, School of Mathematical Sciences, Fudan University, Shanghai 200433, China. The motion of molecular motor is essential to the biophysical functioning of living cells. In principle, this motion can be regraded as a multiple chemi- cal states process. In which, the molecular motor can jump between different chemical states, and in each chemical state, the motor moves forward or back- ward in a corresponding potential. So, mathematically, the motion of molecular motor can be described by several coupled one-dimensional hopping models or by several coupled Fokker-Planck equations. To know the basic properties of molecular motor, in this paper, we will give detailed analysis about the sim- plest cases: in which there are only two chemical states. Actually, many of the existing models, such as the flashing ratchet model, can be regarded as a two-state model. From the explicit expression of the mean velocity, we find that the mean velocity of molecular motor might be nonzero even if the po- tential in each state is periodic, which means that there is no energy input to the molecular motor in each of the two states. At the same time, the mean velocity might be zero even if there is energy input to the molecular motor. Generally, the velocity of molecular motor depends not only on the potentials (or corresponding forward and backward transition rates) in the two states, but also on the transition rates between the two chemical states. PACS numbers: 87.16.Nn, 87.16.A-, 82.39.-k, 05.40.Jc Keywords: molecular motor, one-dimensional hopping model, Fokker-Planck equation ∗Email: [email protected] I. INTRODUCTION 2 Molecular motors are biogenic force generators acting in the nanometer range, and converting chemical energy into mechanical work [1, 2], which play essential roles in eukaryotic cells [3 -- 7]. In the super family of molecular motors [8], the most extensively studied ones are conventional kinesin [9 -- 16], cytoplasmic dynein [17 -- 23], myosin V [24 -- 29], and F0F1−ATPase [30 -- 36]. The conventional kinesin can walk hand-over- hand along microtubule about 1 µm to the plus end direction of the microtubule before its dissociation from the track [37 -- 39], with step size 8.2 nm [40 -- 42] and stall force 6−8 pN [10, 11, 14, 19, 43 -- 47], which is independent of ATP concentration [10]. In saturating ATP solution, its zero load velocity is about 700−1000 nm/s [10, 46, 48]. Cytoplasmic dynein also can walk hand-over-hand along microtubule with average step size 8.2 nm [22, 49 -- 52], but to the minus end direction [18]. Recent experimental data indicate that its stall force is also about 6−8 pN [49, 52, 53], and independent of ATP concentration [49]. To the dynein which is purified from mammalian animals, its maximal velocity is also about 700−1000 nm/s [18, 54, 55]. Myosin V is also a processive motor but walks along actin filaments with average step size 36 nm, and ATP independent stall force 2−3 pN [56 -- 61]. ATPase consists of two portions F0 and F1 connected to a γ shaft. It can use the proton-motive force across the mitochondrial membranes to make ATP from ADP and Pi, and also can use ATP to drive the rotation of the γ shaft [62]. Recent experiments found that there are also many other molecular motors that can move processively, such as kinesin CENP-E [63], myosin VI [64 -- 67], myosin VIIa [68], myosin IXb [69], myosin XI [70], and T7 DNA helicase [71]. There are many mathematical models to describe the motion of molecular motor, such as Fokker-Planck equation [2, 72 -- 74], Langevin equation [75], and master equa- tion [9, 76 -- 79]. However, so far, almost all of the explicit formulations of biophysical properties of molecular motor, such as mean velocity [2, 80], effective diffusion con- stant [81, 82], and mean first passage time [83, 84], are obtained by employing one-sate models, in which the molecular motor moves along its track in one tilted periodic po- tential [108]. One of the basic properties of such models is that the mean velocity of 3 molecular motor does not vanish as long as the input energy is positive. These models and their corresponding results are valuable to describe the tightly mechanochemi- cal coupled cases of motor motion. However, recent experimental data indicate the motion of molecular motors, including conventional kinesin [43, 85 -- 88], cytoplasmic dynein [89], myosin II [90, 91], and F1-ATPase [92] are usually loosely coupled to ATP hydrolysis, i.e., the input energy might be nonzero even if the mean velocity vanishes. To study these loosely coupled cases, it is necessary to use multi-state models. In fact, the multi-state models have been used by some authors [73, 93, 94]. However, it is hard to get meaningful explicit results for the general N-state models. Usually the numerical calculations are employed [95 -- 97]. In this paper, we will give a detailed theoretical analysis to the two-state models. Actually, the two-state models have most of the essential properties of the general multi-state models, and they have been used in many studies [95, 98 -- 104]. There are two different forms of two-state models: (1) two coupled one-dimensional hop- ping models, and (2) two coupled one-dimensional Fokker-Planck equations, which is equivalent to two coupled Langevin equations (in fact, it also can be verified that, any one-dimensional hopping model can be well approximated by a one-dimensional Fokker-Planck equation [105]). In the following, we will give the explicit formulation of the mean velocity of molecular motor by using two coupled one-dimensional hop- ping models and two coupled one-dimensional Fokker-Plank equations respectively. From this formulation, the stall force, i.e., the external load under which the mean velocity vanishes, can be obtained. We find that, the mean velocity, and consequently the stall force depend not only on potentials in the two states (or corresponding for- ward and backward transition rates), but also on transition rates between the two chemical states. In general, part of the input energy will dissipate into the environ- ment, and so the energy efficiency, i.e., the ratio of mechanical work done by the molecular motor to the input energy, might be far less than 1. For example, the mean velocity might be zero even if the input energy in each state is nonzero. The organization of this paper is as follows. In the next section, the two cou- pled one-dimensional hopping models are discussed, and then in Section III, the two coupled Fokker-Planck equations are analyzed. In each models, three special cases 4 are further analyzed: (1) The motor can jump between the two chemical states at only one position. In fact, the properties of this special case are much similar as the usual one state model. At steady state, there is no energy input to the molecular motor during its transition between the two chemical states. (2) The motor can jump between the two states at two positions. This special case has the typical properties of the general cases. (3) One of the two potentials is constant, or all the correspond- ing transition rates (forward and backward) are equal to each other in one of the two states. This special case corresponds to the flashing ratchet model of molecular motors. Finally, the results are briefly summarized in Section IV. II. TWO COUPLED ONE-DIMENSIONAL HOPPING MODELS The two coupled one-dimensional hopping models are schematically depicted in Fig. 1. In which, the forward and backward transition rates in state 1 are denoted by Fn (n → n + 1) and Bn (n → n − 1), the forward and backward transition rates in state 2 are denoted by fn (n → n + 1) and bn (n → n − 1), and the transition rates between the two states at position n are denoted by ωn a (state 1 → state 2) and ωn d (state 2 → state 1). Under the assumption of periodicity, we have FlN +n = Fn, BlN +n = Bn, flN +n = fn, blN +n = bn, ωlN +n a ωlN +n d = ωn d , = ωn d , (1) where l is an integer number l, N is the period of hopping models. Let Pn(t) be the probability of finding molecular motor at position n of state 1 (denoted by 1n) at time t, and ρn(t) be the probability of finding molecular motor at position n of state 2 (denoted by 2n) at time t. Then the evolution of probabilities Pn(t) and ρn(t) are governed by the following master equations: Pn(t) = Fn−1 Pn−1(t) − (Fn + Bn) Pn(t) + Bn+1 Pn+1(t) − ωa n Pn(t) + ωd n ρn(t), ρn(t) = fn−1 ρn−1(t) − (fn + bn)ρn(t) + bn+1 ρn+1(t) + ωa n Pn(t) − ωd n ρn(t), n = 0, ±1, ±2, · · · . (2) d dt d dt   Let ∞ Pn(t) = PlN +n(t), ρn(t) = Xl=−∞ then, at steady state, Pn and ρn satisfy [106] ρlN +n(t), ∞ Xl=−∞ Fn−1Pn−1 − (Fn + Bn)Pn + Bn+1Pn+1 − ωa nPn + ωd nρn = 0, fn−1ρn−1 − (fn + bn)ρn + bn+1ρn+1 + ωa nPn − ωd nρn = 0,   with n = 1, 2, · · · , N, and the total flux of probability, Jn+ 1 2 = (FnPn − Bn+1Pn+1) + (fnρn − bn+1ρn+1), is constant, i.e. Jn+ 1 2 ≡ J for n = 1, 2, · · · , N. From the first equation of (4), one sees that ρn = [(Fn + Bn + ωn a )Pn − Fn−1Pn−1 − Bn+1Pn+1]/ωn d . Substituting (6) into (5), one can easily verify J = An−1Pn−1 + CnPn + Dn+1Pn+1 + En+2Pn+2, where An = −fn+1Fn/ωn+1 d , Cn = [fn(Fn + Bn + ωn a )]/ωn d + Fn + bn+1Fn/ωn+1 , d Dn = −[Bn + fn−1Bn/ωn−1 d + bn(fn + Bn + ωn a )/ωn d ], En = bn−1Bn/ωn−1 d .   By (7) and routine analysis, we obtain Pi = XiJ + YiP1 + ZiP2 + WiP3, where XN = 1 AN YN = − C1 AN , ZN = − D2 AN , WN = − E3 AN , XN −1 = 1 AN −1 (cid:16)1 − CN AN(cid:17) , − E2 AN −1 YN −1 = CN C1 AN −1AN , WN −1 = CN E3 AN −1AN − D1 AN −1 , , ZN −1 = CN D2 AN −1AN 5 (3) (4) (5) (6) (7) (8) (9) (10) (11) XN −2 = − 1 AN −2 E1 CN −1 AN −2AN −1 CN −1D1 YN −2 = − ZN −2 = + AN −2 CN −1E2 AN −2AN −1 WN −2 = DN E3 AN −2AN − AN −2AN −1 DN D2 AN −2AN CN −1CN E3 + − AN −2AN −1AN , + CN −1CN AN −2AN −1AN − DN + DN C1 AN −2AN − AN −2AN CN −1CN C1 AN −2AN −1AN CN −1CN D2 AN −2AN −1AN , 6 , , (12)   and for 1 ≤ k ≤ N − 3, 1 − (Ck+1Xk+1 + Dk+2Xk+2 + Ek+3Xk+3) Xk = Yk = − Ck+1Yk+1 + Dk+2Yk+2 + Ek+3Yk+3 Ak Ak , Zk = − Ck+1Zk+1 + Dk+2Zk+2 + Ek+3Zk+3 Ak , Wk = − Ck+1Wk+1 + Dk+2Wk+2 + Ek+3Wk+3 Ak . , (13)   Then, for i = 1, 2, 3 in equation (9), we obtain the following equations AP = JX, (14) where A =  1 − Y1 −Z1 −W1 −Y2 1 − Z2 −W2 −Y3 −Z3 1 − W3   , P =  P1 P2 P3   , X =  X1 X2 X3   . (15) So, Pi = PiJ for i = 1, 2, 3, with P = ( P1, P2, P3)T satisfy A P = X. Consequently, Pi, for 3 < i ≤ N, can be obtained by Eq. (9), Pi = (Xi + Yi P1 + Zi P2 + Wi P3)J =: PiJ, (16) and therefore, ρi can be obtained by Eq. (6), ρi =[(Fi + Bi + ωi =[(Fi + Bi + ωi a)Pi − Fi−1Pi−1 − Bi+1Pi+1]/ωi d, a) Pi − Fi−1 Pi−1 − Bi+1 Pi+1]J/ωi d, (17) = : ρiJ. The probability flux J in Eqs. (16) (17) is determined by the normalization condition N (Pi + ρi) = 1, Pi=1 N 7 ( Pi + ρi), J =1/ =1/ N Xi=1 Xi=1 (cid:20) ωi d a + ωi ωi d +(cid:18) 1 ωi d − 1 d (cid:19) Fi +(cid:18) 1 ωi d ωi+1 − 1 d (cid:19) Bi(cid:21) Pi. ωi−1 Specially, if ωi a ≡ ωa and ωi d ≡ ωd are constants, then J = ωd (ωa + ωd) = Pi N Pi=1 (ωa + ωd) ωd (Xi + Yi P1 + Zi P2 + Wi P3) N Pi=1 A. Special case I: ωi a = ωi d = 0 for 1 ≤ i ≤ N − 1 (18) . (19) For convenience, we denote ωN a , ωN d by ωa, ωd respectively (see Fig. 2). For this special case, the steady state probabilities Pn, ρn satisfy FN PN − B1P1 = F1P1 − B2P2 = · · · = FN −1PN −1 − BN PN =: J, fN ρN − b1ρ1 = f1ρ1 − b2ρ2 = · · · = fN −1ρN −1 − bN ρN =: j.   It can be readily verified that Pk = XkPN − YkJ, k 1 Fk Fi−1 Bi , Yk = Xk = Yi=1 PN = N Bi ! PN − 1 Yi=1 Fi−1 FN k k Xi=1 Yj=i Fj Bj . N N Xi=1 Yj=i Fj Bj! J, with Specially, which implies Fj Bj J. N N Pi=1 Qj=i Fi−1 Bi − 1 PN = 1 FN N Qi=1 (20) (21) (22) (23) (24) Combining (21) (22) and (24), one finds Fj Bj!(cid:18) k Qi=1 − 1 1 N N   N Pi=1 FN N Pi=1 Qj=i Qi=1 Bj!(cid:18) k Qi=1 Qj=i Fi−1 Bi Fj N Pk = = Fi−1 Bi (cid:19) − k Qj=i − 1 Fi Bi(cid:19) − k Pi=1 Qi=1 Fi−1 Bi N k k 1 Fk Xi=1 Yj=i Bj!(cid:18) N Qi=1 Fj Fj Bj Fi−1 Bi Using the periodic conditions (1), one can verify that 1 Fk Pk = N +k N +k Fj Bj J. Qj=i − 1 1 fk ρk = N +k N +k fj bj j. Qj=i − 1 N Pi=k+1 Qi=1 Fi Bi N Pi=k+1 Qi=1 fi bi Using the same method, the probability ρk can be obtained At steady state, ωaPN = ωdρN , which implies J,   − 1(cid:19) J Fk . 8 (25) (26) (27) j = 1 FN 1 fN N N N Pi=1 Pi=1 N Qj=i Qj=i Therefore, Fj Bj (cid:30)(cid:18) N Qi=1 bj (cid:30)(cid:18) N Qi=1 fj Fi Bi fi bi − 1(cid:19) − 1(cid:19) ωa ωd J =: ΞJ. (28) N +k Qj=i − 1 ρk = 1 fk N N +k N Pi=k+1 Qi=1 Fi Bi fj bj ΘJ, where Θ = ωa FN ωd fN Fj Bj fj bj . N N Pi=1 Pi=1 N N Qj=i Qj=i (29) (Pk + ρk) = 1, from (26, 29), the probability flux J can be Since Pk, ρk satisfy obtained as follows Pk+1 φ(cid:18) N Qi=1 Fi Bi − 1(cid:19) φΨ + Φψ J = . (30) where φ = ωd fN Φ = ωa FN fj bj , Fj Bj , ψ = Ψ = N N N N Pi=1 Qj=i Qj=i Pi=1 So the total flux of this system is J + j = (1 + Ξ)J =  1 + N N Qi=1 Qi=1 fi bi − 1 Fi Bi − 1 Θ  N +k N +k Qj=i Qj=i N N fk Fk N +k N +k Pi=k+1 Pi=k+1 Pk=1 1 Pk=1 1 φ(cid:18) N Qi=1 J = 9 (31) fj bj! , Bj! . Fj Fi Bi − 1(cid:19) + Φ(cid:18) N Qi=1 φΨ + Φψ fi bi − 1(cid:19) . (32) Combining (26) (29) and (30), the probabilities Pk and ρk can be obtained as follows N +k N +k Xi=k+1 Yj=i Qi=1 N fi bi Fj Bj , = N Qi=1 Pk = φ φΨ + Φψ 1 Fk ρk = Φ φΨ + Φψ 1 fk N +k N +k Xi=k+1 Yj=i fj bj . (33) By (32), one easily sees that, if Fi Bi = 1, then J = j = 0, consequently the total probability flux J + j = 0. In other words, for this special case, if there is no energy input to the molecular motor in each state, then the mean velocity would be zero. But the reverse does not hold. Note, the potential changes in one period of state 1 and state 2 are ∆G1 = kBT ln(cid:16)QN respectively [9, 107]. i=1 Fi Bi(cid:17) and ∆G2 = kBT ln(cid:16)QN i=1 fi bi(cid:17) B. Special case II: ωi a = ωi d = 0 for i 6= M, N For convenience, we denote ωN a , ωN d by Ωa, Ωd, and ωM a , ωM d by ωa, ωd (see Fig. 3). At steady state, Pk, ρk satisfy 10   FN PN − B1P1 = F1P1 − B2P2 = · · · = FM −1PM −1 − BM PM =: J1, FM PM − BM +1PM +1 = · · · = FN −1PN −1 − BN PN =: J2, fN ρN − b1ρ1 = f1ρ1 − b2ρ2 = · · · = fM −1ρM −1 − bM ρM =: j1, fM ρM − bM +1ρM +1 = · · · = fN −1ρN −1 − bN ρN =: j2, ωaPM + ΩaPN = ωdρM + ΩdρN , J2 = J1 + ΩaPN − ΩdρN , j2 = j1 − ΩaPN + ΩdρN , N (34) (Pk + ρk) = 1. Xk=1 From the first equation in (34), one can easily get Pk = k Yi=1 Fi−1 Bi ! PN − 1 Fk k Xi=1 k Yj=i Fj Bj! J1, 1 ≤ k ≤ M. (35) At the same time, from the second equation in (34), 1 1 1 − k Fj Fi−1 Fi−1 Fi−1 Pk = k Fk k Bj! J2 Bi ! PM − Yi=M +1 Xi=M +1 Yj=i = k FM M Bi ! PN − Bi !" M Yi=M +1 Xi=1 Yi=1 Bj! (J1 + ΩaPN − ΩdρN ) Fk k Yj=i Xi=M +1 =" k Fk k Yi=1 Yj=i Xi=M +1 Bj! ρN − Fk k Yj=i Xi=M +1 Bj!# PN Fk k Xi=1 Fi−1 Bi Yj=i Ωa Ωd Fj Fj k Fj − + k 1 k k Fj Bj! J1 Fj Bj! J1# M Yj=i (36) = : (Rk − ΩaSk)PN + ΩdSkρN − TkJ1, for M + 1 ≤ k ≤ N. In particular, PN = (RN − ΩaSN )PN + ΩdSN ρN − TN J1, which gives J1 = (RN − ΩaSN − 1)PN + ΩdSN ρN TN . (37) Substituting (37) into (35) (36), we obtain Pk = GkPN + HkρN , where Similarly, Gk =  Hk =  Rk − Tk TN Rk − ΩaSk − Tk TN (RN − ΩaSN − 1), 1 ≤ k ≤ M, (RN − ΩaSN − 1), M + 1 ≤ k ≤ N, − Tk TN − Tk TN ΩdSN , 1 ≤ k ≤ M, ΩdSN + ΩdSk, M + 1 ≤ k ≤ N. ρk = gkρN + hkPN , 11 (38) (39) (40) where gk, hk, and the corresponding rk, sk, tk in expressions of gk, hk, can be obtained by replacing Fj, Bj, Ωa, Ωd in the expressions of Rk, Sk, Tk, Gk, Hk with fj, bj, Ωd, Ωa respectively. Combining (38) (40) and the fifth equality in (34), we have ωa(GM PN + HM ρN ) + ΩaPN = ωd(gM ρN + hM PN ) + ΩdρN , (41) i.e. So (Ωa + ωaGM − ωdhM )PN = (Ωd + ωdgM − ωaHM )ρN . (42) ρN = Ωa + ωaGM − ωdhM Ωd + ωdgM − ωaHM PN =: U V PN . From (38) (40) and (43), one finds Pk =(cid:18)Gk + U V Hk(cid:19) PN , ρk =(cid:18)hk + U V gk(cid:19) PN . In view of the last equation in (34), one gets " N Xk=1(cid:18)(Gk + hk) + U V (gk + Hk)(cid:19)# PN = 1, which implies PN = 1 . N Pk=1(cid:2)(Gk + hk) + U V (gk + Hk)(cid:3) (43) (44) (45) (46) 12 (47) U V . . (48) By (37) (43) (46), we have J1 = TN Similarly, one can verify that j1 = tN (RN − ΩaSN − 1) + ΩdSN N Pk=1(cid:2)(Gk + hk) + U V (gk + Hk)(cid:3) (rN − ΩdsN − 1) U N V + ΩasN Pk=1(cid:2)(Gk + hk) + U V (gk + Hk)(cid:3) Therefore, the total flux of this special case is J1 + j1 = (cid:0)(RN − ΩaSN − 1) + ΩdSN N 1 TN tN Pk=1(cid:2)(Gk + hk) + U (cid:18)sN Pk=1(cid:2)(Gk + hk) + U [(Gk + hk)V + (gk + Hk)U](cid:18) sN Pk=1 V (gk + Hk)(cid:3) ΩaωdrM − ωaΩdRM tN tN N N − J1 + j1 = = where More specially, if RN = rN = 1, then the total probability flux is U =Ωa + ωaRM + ωaΩaSN TM /TN + ωdΩasN tM /tN > 0, V =Ωd + ωdrM + ωdΩdsN tM /tN + ωaΩdSN TM /TN > 0. So the direction of probability flux is determined by the sign of (cid:16) sN (ΩaωdrM − ωaΩdRM ). One can see that, RN = rN = 1, i.e., ∆G1 = ∆G2 = 0, does TN(cid:17) and − SN tN not read the mean velocity vanishes. To better understand the inter-state transition rates dependence of the total prob- ability flux, we assume that (Ωa, Ωd, ωa, ωd) = λ( Ωa, Ωd, ωa, ωd). It can be verified that the total probability flux J := J1 + j1 in (50) increases mono- tonically with parameter λ. If λ = 0 then J = 0. If λ → ∞, them J tends to Ωa ωdrM − ωa ΩdRM ∗ (cid:18)sN tN − SN TN(cid:19) , (52) U V(cid:1) tN +(cid:0)(rN − ΩdsN − 1) U V (gk + Hk)(cid:3) V + ΩasN(cid:1) TN U V Ωd(cid:19) , SN TN(cid:19)(cid:18)Ωa − TN(cid:19) , SN − . (49) (50) (51) where 13 )Rk + ( ωa ΩaSN TM TN + N tN ωd ΩasN tM )rk# (sk − Sk)# − tksN tN (cid:19) + Xk=M +1 N ∗ = + tN TN ωd ΩdsN tM ωa ΩdSN TM Xk=1"( + ( Ωa ωdrM − ωa ΩdRM )" N Xk=1(cid:18) TkSN TN (cid:19) N =(cid:18) ωdsN tM Xk=1 + ( Ωa ωdrM − ωa ΩdRM )" N Xk=1(cid:18) TkSN ωaSN TM TN TN + tN ( ΩdRk + Ωark) − tksN tN (cid:19) + N Xk=M +1 (sk − Sk)# . C. Special case III: ωi a = ωi d = 0 for i 6= M, N , and fi = bi ≡ f for 1 ≤ i ≤ N As pointed out in the Introduction, the flashing ratchet model can be regarded as one example of this special case. For this more special case, we have rk ≡ 1, tk = k f , for 1 ≤ k ≤ N, (53) and sk = (k − M)/f for M + 1 ≤ k ≤ N. It can be easily verified that 1 + (N −M )k 1 + M (N −k) N f Ωd, N f Ωd, − (N −M )k − M (N −k) N f Ωa, N f Ωa, gk =  hk =  1 ≤ k ≤ M, M + 1 ≤ k ≤ N, 1 ≤ k ≤ M, M + 1 ≤ k ≤ N. So U =Ωa + ωaGM − ωdhM =Ωa + ωa(cid:18)RM + TM TN V =Ωd + ωdgM − ωaHM SN Ωa − TM TN (RN − 1)(cid:19) + M(N − M) Nf ωdΩa, =Ωd + ωd(cid:18)1 + M(N − M) Nf Ωd(cid:19) + TM TN SN ωaΩd. Moreover, if RN = 1 then the total probability flux is J1 + j1 = Ωaωd − ωaΩdRM ∆ (cid:18)N − M N − SN TN(cid:19) , (54) (55) (56) where M ∆ = (V Rk + Urk) + (ΩaωdrM − ωaΩdRM )" M Xk=1(cid:18)TkSN TN Xk=1 =(Ωaωd − ωaΩdRM )"SN TN Tk − M Xk=1 N Xk=M +1 Sk + (N − M)[(N − M)(N + M + 1) − MN] 2Nf # 14 − tksN tN (cid:19) + N Xk=M +1 (sk − Sk)# , + MU + V Rk. M Xk=1 III. TWO COUPLED FOKKER-PLANCK EQUATIONS The general two coupled Fokker-Planck equations are as follows   ∂t P =D∂x(βV ′ 1 P + ∂x P ) + ωd(x)ρ − ωa(x) P , ∂t ρ =D∂x(βV ′ 2 ρ + ∂x ρ) − ωd(x)ρ + ωa(x) P , − ∞ ≤ x ≤ +∞, (57) where D is free diffusion constant, β = 1/kBT with kB is Boltzmann constant, and T is absolute temperature, P (x, t) and ρ(x, t) are probability densities of finding molecular motor at position x at time t and in states 1 and 2 respectively, V1, V2 are (tilted) periodic potentials with period L. ωa(x), ωd(x) are transition rates between states 1 and 2 at position x [109]. Similar as in [82], let P (x, t) = +∞ Xk=−∞ then P (x, t), ρ(x, t) satisfy P (x + kL, t), ρ(x, t) = +∞ Xk=−∞ ρ(x + kL, t), (58)   ∂tP =D∂x(βV ′ 1P + ∂xP ) + ωd(x)ρ − ωa(x)P, ∂tρ =D∂x(βV ′ 2ρ + ∂xρ) − ωd(x)ρ + ωa(x)P, 0 ≤ x ≤ L. (59) The steady state solution of (59) can be obtained under the following constraints: P (0) = P (L), ρ(0) = ρ(L), (P + ρ)dx = 1, Z L 0 Z L 0 ωdρdx =Z L 0 ωaP dx. (60) The corresponding probability flux is J = −D (βV ′ 1P + ∂xP + βV ′ 2ρ + ∂xρ) , (61) 15 and the mean velocity of molecular motor is V =R L 0 Jdx = −βDR L JL. If ωa(x), ωd(x) are constants, Eq. (59) had been discussed by Y.-D. Chen [95], 0 (V ′ 1P + V ′ 2ρ)dx = and it can be solved numerically using the similar method as the one used in WPE method [96, 97]. A. Special case I: ωa(x) = ωd(x) ≡ 0 for 0 < x < L For this special case, the steady state probability densities P (x), ρ(x) of finding molecular motor at position x are governed by the following equations D∂x(βV ′ 1P + ∂xP ) = 0, D∂x(βV ′ 2ρ + ∂xρ) = 0, 0 < x < L. (62) Meanwhile, P (x), ρ(x) satisfy the following boundary conditions and normalization constraint: P (0) = P (L), ρ(0) = ρ(L), ωaP (0) = ωdρ(0), (P + ρ)dx = 1, (63) Z L 0 where ωa = ωa(L), ωd = ωd(L). The probability fluxes in the two states are J = −D(βV ′ 1P + ∂xP ), j = −D(βV ′ 2ρ + ∂xρ). (64) So Eqs. (62) can be reformulated as     βV ′ 1P + ∂xP = −J/D, βV ′ 2ρ + ∂xρ = −j/D, 0 < x < L. (65) The general solutions of (65) are P (x) =(cid:18)− J D Z x 0 eβV1(y)dy + C1(cid:19) e−βV1(x), ρ(x) =(cid:18)− j D Z x 0 eβV2(y)dy + C2(cid:19) e−βV2(x), where the constants C1, C2 can be determined by the periodic boundary conditions P (0) = P (L), ρ(0) = ρ(L): C1 = 0 eβV1(y)dy 1 − e−β∆V1 J DR L with ∆Vi = Vi(0) − Vi(L). Therefore P (x) = J x DR x+L eβ[V1(y)−V1(x)]dy 1 − e−β∆V1 , , C2 = 0 eβV2(y)dy 1 − e−β∆V2 J DR L , ρ(x) = j x DR x+L eβ[V2(y)−V2(x)]dy 1 − e−β∆V2 . (66) 16 , j 0 eβ[V2(y)−V2(0)]dy 1 − e−β∆V2 From ωaP (0) = ωdρ(0), one sees J so ωa j = = ωd DR L 0 eβ[V1(y)−V1(0)]dy 1 − e−β∆V1 DR L ωa(cid:0)eβV2(0) − (eβV2(L)(cid:1)R L ωd (eβV1(0) − (eβV1(L))R L From (66) (67) and the normalization condition R L ωdD(cid:0)eβV1(0) − eβV1(L)(cid:1)R L ωaD(cid:0)eβV2(0) − eβV2(L)(cid:1)R L J = j = ⋆ ⋆ 0 eβV1(y)dy 0 eβV2(y)dy J. (67) 0 (P + ρ)dx = 1, one can easily get 0 eβV2(y)dy 0 eβV1(y)dy , , (68) where ⋆ =ωdeβV1(0)(cid:18)Z L + ωaeβV2(0)(cid:18)Z L eβV2(y)dy(cid:19)(cid:20)Z L eβV1(y)dy(cid:19)(cid:20)Z L e−βV1(x)(cid:18)Z x+L e−βV2(x)(cid:18)Z x+L eβV1(y)dy(cid:19) dx(cid:21) eβV2(y)dy(cid:19) dx(cid:21) . x x 0 0 0 0 It can be easily found that, for this special case, the total probability flux J + j = 0 if potentials V1, V2 are periodic, i.e., ∆V1 = ∆V2 = 0. From (32) and (68), one sees that, the properties of this special case are similar as those of the special case I of the two coupled one-dimensional hopping models [105]. B. Special case II: ωa(x) = ωd(x) ≡ 0 for x 6= a, L For this special case, the governing equations of the steady state probability den- sities P (x), ρ(x) are D∂x(βV ′ 1P1 + ∂xP1) = 0, D∂x(βV ′ 2ρ1 + ∂xρ1) = 0, D∂x(βV ′ 1P2 + ∂xP2) = 0, D∂x(βV ′ 2ρ2 + ∂xρ2) = 0, 0 < x < a, a < x < L,     (69) with the following constraints P1(0) = P2(L), P1(a) = P2(a), ρ1(0) = ρ2(L), ρ1(a) = ρ2(a), J1 = J2 + ωaP (a) − ωdρ(a), j1 = j2 − ωaP (a) + ωdρ(a), ωaP (a) + ΩaP (L) = ωdρ(a) + Ωdρ(L), Z a 0 (P1 + ρ1)dx +Z L a (P2 + ρ2)dx = 1, 17 (70) where ωa = ωa(a), ωd = ωd(a), Ωa = ωa(L), Ωd = ωd(L), and Ji = −D(βV ′ 1Pi + ∂xPi) ji = D − (βV ′ 2ρi + ∂xρi), for i = 1, 2, are probability fluxes in the two states. The general solutions of (69) can be written as follows Pi(x) = −Fi(x)Ji + Gi(x)Ci, ρi(x) = −fi(x)ji + gi(x)ci, i = 1, 2, (71) where F1(x) = 1 F2(x) = 1 f1(x) = 1 f2(x) = 1 D e−βV1(x)R x D e−βV1(x)R x D e−βV2(x)R x D e−βV2(x)R x 0 eβV1(y)dy, G1(x) = e−βV1(x), a eβV1(y)dy, G2(x) = e−βV1(x), g1(x) = e−βV2(x), 0 eβV2(y)dy, a eβV2(y)dy, g2(x) = e−βV2(x), 0 ≤ x ≤ a, a ≤ x ≤ L, 0 ≤ x ≤ a, a ≤ x ≤ L. (72) From (70) (71), one can verify that Ji, ji and Ci, ci satisfy the following equations 18 G1(0)C1 = −F2(L)J2 + G2(L)C2, − F1(a)J1 + G1(a)C1 = G2(a)C2, g1(0)c1 = −f2(L)j2 + g2(L)c2, − f1(a)j1 + g1(a)c1 = g2(a)c2, J1 = J2 + ωaG2(a)C2 − ωdg2(a)c2, J1 + j1 = J2 + j2, ωaG2(a)C2 + ΩaG1(0)C1 = ωdg(a)c2 + Ωdg1(0)c1, (73) 0 −(cid:18)Z a −(cid:18)Z a 0 F1dx(cid:19) J1 +(cid:18)Z a f1dx(cid:19) j1 +(cid:18)Z a G1dx(cid:19) C1 −(cid:18)Z L g1dx(cid:19) c1 −(cid:18)Z L F2dx(cid:19) J2 +(cid:18)Z L f2dx(cid:19) j2 +(cid:18)Z L G2dx(cid:19) C2, g2dx(cid:19) c2 = 1. 0 0 a a a a For the sake of convenience, we rewrite equations in (73) as AX = B, with X = (J1, C1, J2, C2, j1, c1, j2, c2)T , B = (0, 0, 0, 0, 0, 0, 0, 1)T and A = −F1(a) G1(a) 0 0 1 1 0 0 0 0   ,   0 G1(0) F2(L) −G2(L) 0 0 0 −G2(a) 0 0 0 0 0 0 0 0 0 0 0 g1(0) f2(L) −g2(L) −f1(a) g1(a) 0 −g2(a) −1 −ωaG2(a) −1 0 0 1 0 0 0 ωdg2(a) −1 0 −IF1 IG1 −IF2 IG2 −If1 Ig1 −If2 Ig2 in which IH1 = R a a H2dx for H = F, G, f, g. Although it can be obtained explicitly, the solution of AX = B is very complex. So, for simplicity, we 0 H1dx, IH2 = R L only discuss the special cases in which potentials V1, V2 satisfy ∆V1 = ∆V2 = 0. By routine analysis, one can obtain J1 = F2(L)G1(a)[−ωaΩdG1(a)g1(0)f2(L)g1(a) − ωaΩdG1(a)f1(a)g1(0)2 ΩaωdG1(0)g1(a)2f2(L) + ΩaωdG1(0)f1(a)g1(0)g1(a)]/ det(A), J2 = −F1(a)G1(0)[−ωaΩdG1(a)g1(0)f2(L)g1(a) − ωaΩdG1(a)f1(a)g1(0)2 ΩaωdG1(0)g1(a)2f2(L) + ΩaωdG1(0)f1(a)g1(0)g1(a)]/ det(A), 19 j1 = −f2(L)g1(a)[−ωaΩdF1(a)G1(0)G1(a)g1(0) − ωaΩdF2(L)G1(a)2g1(0) +ΩaωdF1(a)G1(0)2g1(a) + ΩaωdG1(0)F2(L)G1(a)g1(a)]/ det(A), j2 = f1(a)g1(0)[−ωaΩdF1(a)G1(0)G1(a)g1(0) − ωaΩdF2(L)G1(a)2g1(0) +ΩaωdF1(a)G1(0)2g1(a) + ΩaωdG1(0)F2(L)G1(a)g1(a)]/ det(A), where det(A) is the determinant of matrix A, and it can be proved that det(A) < 0. So the total probability flux is J1 + j1 =J2 + j2 =[ΩaωdG1(0)g1(a) − ωaΩdG1(a)g1(0)] × [f1(a)g1(0)F2(L)G1(a) − F1(a)G1(0)f2(L)g1(a)]/ det(A) [g1(0)G1(0)]2f1(a)F1(a) G1(a) Obviously, Ji + ji > 0 if and only if (cid:2)Ωaωdeβ(V2(0)−V2(a)) − ωaΩdeβ(V1(0)−V1(a))(cid:3) × 0 eβV1(y) − R L h R L the properties of this special case are similar as those of the special case II of the cou- In view of the expression in (50), one can find that, a eβV1(y) R a pled one-dimensional hopping models. The mean velocity of molecular motor might not be zero even if there is no energy input in each state. For this special case, the energy for motor motion comes from the processes that drive the motor from one state to another [98 -- 101]. g1(a) − ωaΩd g1(a) g1(0) G1(0)(cid:21) (cid:20)Ωaωd g1(0)(cid:21) (cid:2)Ωaωdeβ(V2(0)−V2(a)) − ωaΩdeβ(V1(0)−V1(a))(cid:3) (74) = = det(A) G1(a) G1(0) ×(cid:20) F2(L) f2(L) f1(a) [g1(0)G1(0)]2f1(a)F1(a) F1(a) − a eβV2(y) 0 eβV2(y)# . − R L R a det(A) a eβV1(y) 0 eβV1(y) ×"R L R a 0 eβV2(y)i < 0. a eβV2(y) R a C. Special case III: ωa(x) = ωd(x) ≡ 0 for x 6= a, L, and V2(x) is constant For this special case, the governing equations of steady state probability densities P (x), ρ(x) are as follows 20 D∂x(βV ′ 1P1 + ∂xP1) = 0, D∂2 xρ1 = 0, D∂x(βV ′ 1P2 + ∂xP2) = 0, D∂2 xρ2 = 0, 0 < x < a, a < x < L. (75)     Its general solutions are (71) but with fi(x) = x/D, gi(x) ≡ 1. The solution which satisfies the constraints (70) is as follows J1 = −2[ωdG1(0)G2(a)DL − ωdG1(a)G2(L)DL + ΩdG1(0)G2(a)DL −ΩdG1(a)G2(L)DL + ΩdωdG1(0)G2(a)aL − ΩdωdG1(0)G2(a)a2 −ΩdωdG1(a)G2(L)aL + ΩdωdG1(a)G2(L)a2 −ωaΩdG1(a)F2(L)G2(a)DL + ΩaωdG1(0)F2(L)G2(a)LD]/ det(A), J2 = −2[ωaΩdF1(a)G1(0)LDG2(a) − ΩaωdF1(a)G1(0)LDG2(L) +ωdG1(0)G2(a)DL − ωdG1(a)G2(L)DL + ΩdG1(0)G2(a)DL −ΩdG1(a)G2(L)DL + ΩdωdG1(0)G2(a)aL −ΩdωdG1(a)G2(L)aL − ΩdωdG1(0)G2(a)a2 + ΩdωdG1(a)G2(L)a2]/ det(A), j1 = 2(L − a)[−ωaΩdF1(a)G1(0)G2(a) − ωaΩdG1(a)F2(L)G2(a) +ΩaωdF1(a)G1(0)G2(L) + ΩaωdG1(0)F2(L)G2(a)]D/ det(A), j2 = −2a[−ωaΩdF1(a)G1(0)G2(a) − ωaΩdG1(a)F2(L)G2(a) +ΩaωdF1(a)G1(0)G2(L) + ΩaωdG1(0)F2(L)G2(a)]D/ det(A). So the total probability flux is J1 + j1 =J2 + j2 =2{[G1(a)G2(L) − G1(0)G2(a)][ωdDL + ΩdDL + ωdΩd(L − a)a] (76) + 2aDF2(L)G2(a)[ωaΩdG1(a) − ΩaωdG1(0)] + 2D(L − a)F1(a)G1(0)[ΩaωdG2(L) − ωaΩdG2(a)]}/ det(A). More specially, if potential V1(x) is periodic and continuous at a, then G1(0) = G2(L), G1(a) = G2(a). So 21 [ωaΩdG1(a) − ΩaωdG1(0)][aF2(L)G2(a) − (L − a)F1(a)G1(0)] 2D det(A) 2D(G1(0))2G2(a) det(A) [ωaΩdeβ(V1(0)−V1(a)) − Ωaωd] ×(cid:18)aZ L a eβV1(x)dx − (L − a)Z a 0 eβV1(x)dx(cid:19) [ωaΩdeβ(V1(0)−V1(a)) − Ωaωd] 2D(G1(0))2G2(a) det(A) J1 + j1 = = = = eβV1(x)dx(cid:19) 0 0 eβV1(x)dx − LZ a ×(cid:18)aZ L 2LD(G1(0))2G2(a)R L × L − a det(A) L 0 eβV1(x)dx [Ωaωd − ωaΩdeβ(V1(0)−V1(a))] Therefore, the total probability flux J1 + j1 > 0 if and only if [ωaΩdeβ(V1(0)−V1(a)) − a eβV1(x)dx 0 eβV1(x)dx! − R L R L 0 eβV1(x)dx(cid:17) < 0. Similar as before, from (56) and (77) 0 eβV1(x)dx − LR a (77) one can find the similarity between them [105]. Ωaωd](cid:16)aR L To better understand the properties of our model, we discuss the direction of prob- ability fluxes here. For the special case in which there are only two locations at which the inter-state transition rates are nonzero, i.e., the special case II, there are alto- gether 18 different types of probability flux. Since the states 1 and 2 are temporally symmetric, we restrict our discussion only on the cases in which ωaP (a) − ωdρ(a) ≥ 0 for the continuous model, or ωaPM − ωdρM ≥ 0 for the hopping model. Then, there are altogether 9 different types of probability flux (see Fig. 4). Furthermore, if ∆V1 = ∆V2 = 0, then there is only one type (see the figure (2, 2) in Fig. 4). On the other hand, if the potential V2 is constant (or fi = bi ≡ f for the hopping model), then there are altogether 3 different types of probability flux (see the second column in Fig. 4). IV. CONCLUSIONS 22 In conclusion, two chemical states models of molecular motor are discussed in this paper. For some special cases, explicit expressions of mean velocity are obtained. We find that the mean velocity of molecular motor might not be zero even if both of the potentials are periodic, which means there is no energy input to the molecular motor in each of the chemical states. The energy for the motion molecular motor motion comes from the processes that drive the motor from one state to another. For motor proteins, these processes are ATP hydrolysis. At the same time, from the expression of mean velocity, we find that the velocity of molecular motor might be zero even if there exists nonzero input energy. Which implies that the motion of motor protein is usually loosely coupled to ATP hydrolysis [43, 85 -- 92]. Acknowledgments The author has been funded by the National Natural Science Foundation of China (under Grant No. 10701029). He is grateful to the China Scholarship Council for their financial support of his study in United States and thanks Professor Michael E Fisher, Devarajan Thirumalai, and the Institute for Physical Science and Technology at the University of Maryland for their hospitality. [1] D. Bray, Cell movements: from molecules to motility, 2nd Edn (Garland, New York, 2001). [2] J. Howard, Mechanics of Motor Proteins and the Cytoskeleton (Sinauer Associates and Sunderland, MA, 2001). [3] M. Badoual, F. Julicher, and J. Prost, Proc. Natl. Acad. Sci. USA 99, 6696 (2002). [4] S. Klumpp and R. Lipowsky, Proc. Natl. Acad. Sci. USA 102, 17284 (2005). [5] I. H. Riedel-Kruse, A. Hilfinger, J. Howard, and F. Julicher, HFSP J. 1, 192 (2007). [6] Y. Zhang, Phys. Rev. E 79, 061918 (2009). [7] J. Howard, Annu. Rev. Biophys. 38, 217 (2009). 23 [8] R. D. Vale, Cell 112, 467 (2003). [9] M. E. Fisher and A. B. Kolomeisky, Proc. Natl. Acad. Sci. USA 98, 7748 (2001). [10] N. J. Carter and R. A. Cross, Nature 435, 308 (2005). [11] S. M. Block, Biophys. J. 92, 2986 (2007). [12] Y. Zhang, Biophys. Chem. 136, 19 (2008). [13] E. Toprak, A. Yildiz, M. T. Hoffman, S. S. Rosenfeld, and P. R. Selvin, Proc. Natl. Acad. Sci. USA 106, 12717 (2009). [14] N. R. Guydosh and S. M. Block, Nature 08259 (2009). [15] C. Hyeon, S. Klumppb, and J. N. Onuchic, Phys. Chem. Chem. Phys 11, 4899 (2009). [16] V. Hariharan and W. O. Hancock, Cell. Mol. Bioe. 2, 177 (2009). [17] S. L. Reck-Peterson, A. Yildiz, A. P. Carter, A. Gennerich, N. Zhang, and R. D. Vale, Cell 126, 335 (2006). [18] S. Toba, T. M. Watanabe, L. Yamaguchi-Okimoto, Y. Y. Toyoshima, and H. Higuchi, Proc. Natl. Acad. Sci. USA 103, 5741 (2006). [19] A. Gennerich and R. D. Vale, Curr. Opin. Cell. Biol. 21, 59 (2009). [20] A. Houdusse and A. P. Carter, Cell 136, 395 (2009). [21] A. J. Roberts, N. Numata, M. L. Walker, Y. S. Kato, B. Malkova, T. Kon, R. Ohkura, F. Arisaka, P. J. Knight, K. Sutoh, et al., Cell 136, 485 (2009). [22] J. R. Kardon, S. L. Reck-Peterson, and R. D. Vale, Proc. Natl. Acad. Sci. USA 106, 5669 (2009). [23] A. W. R. Serohijos, W. D. Tsygankov, S. Liu, T. C. Elstonbd, and N. V. Dokholyanz, Phys. Chem. Chem. Phys. 11, 4840 (2009). [24] S. S. Rosenfeld and H. L. Sweeney, J. Biol. Chem. 279, 40100 (2004). [25] T. J. Purcell, H. L. Sweeney, and J. A. Spudich, Proc. Natl. Acad. Sci. USA 102, 13873 (2005). [26] C. Veigel, J. E. Molloy, S. Schmitz, and J. Kendrick-Jones, Nat. Cell. Biol. 5, 980 (2003). [27] T. Sakamoto, M. R. Webb, E. Forgacs, H. D. White, and J. R. Sellers, Nature 455, 128 (2008). [28] J. Del R. Jackson and J. E. Baker, Phys. Chem. Chem. Phys. 11, 4808 (2009). 24 [29] R. Fedorov, M. Bohl, G. Tsiavaliaris, F. K. Hartmann, M. H. Taft, P. Baruch, B. Bren- ner, R. Martin, H.-J. Knolker, H. O. Gutzeit, et al., Nat. Struct. Mol. Biol. 16, 80 (2009). [30] H. Wang and G. Oster, Nature 396, 279 (1998). [31] K. Kinosita, H.Noji, K.Adachi, and R. Yasuda, Phil. Trans. R. Soc. B 355, 473 (2000). [32] T. Nishizaka, K. Oiwa, H. Noji, S. Kimura, E. Muneyuki, M. Yoshida, and K. K. Jr, Nat. Struct. Mol. Biol. 11, 142 (2004). [33] K. Adachi, K. Oiwa, T. Nishizaka, S. Furuike, H. Noji, H. Itoh, M. Yoshida, and J. K. Kinosita, Cell 130, 309 (2007). [34] E. Muneyuki, T. Watanabe-Nakayama, T. Suzuki, M. Yoshida, T. Nishizaka, and H. Noji, Biophys. J. 92, 1806 (2007). [35] W. Junge, H. Sielaff, and S. Engelbrecht, Nature 459, 364 (2009). [36] J. H. M. Jr., V. Vajrala, H. L. Infante, J. R.Claycomb, A. Palanisami, J. Fang, and G. T. Mercier, Physica B 404, 503 (2009). [37] S. M. Block, L. S. B. Goldstein, and B. J. Schnapp, Nature 348, 348 (1990). [38] A. Yildiz, M. Tomishige, R. D. Vale, and P. R. Selvin, Science 303, 676 (2004). [39] C. L. Asbury, A. N. Fehr, and S. M. Block, Science 302, 2130 (2003). [40] M. J. Schnitzer and S. M. Block, Nature 388, 386 (1997). [41] D. L. Coy, M. Wagenbach, and J. Howard, J. Biol. Chem. 274, 3667 (1999). [42] A. N. Fehr, C. L. Asbury, and S. M. Block, Biophys. J. (2007). [43] A. Yildiz, M. Tomishige, A. Gennerich, and R. D. Vale, Cell 134, 1030 (2008). [44] D. D. Hackney, Proc. Natl. Acad. Sci. USA 102, 18338 (2005). [45] Y. Taniguchi, M. Nishiyama, Y. Ishhi, and T. Yanagida, Nat. Chem. Biol. 1, 342 (2005). [46] M. Nishiyama, H. Higuchi, and T. Yanagida, Nature Cell Biol. 4, 790 (2002). [47] M. J. Schnitzer, K. Visscher, and S. M. Block, Nat. Cell. Biol. 2, 718 (2000). [48] S. M. Block, C. L. Asbury, J. W. Shaevitz, and M. J. Lang, Proc. Natl. Acad. Sci. USA 100, 2351 (2003). [49] A. Gennerich, A. P. Carter, S. L. Reck-Peterson, and R. D. Vale, Cell 131, 952 (2007). [50] T. M. Watanabe and H. Higuchi, Biophys. J. 92, 4109 (2007). 25 [51] R. Mallik, B. C. Carter, S. A. Lex, S. J. King, and S. P. Gross, Nature 427, 649 (2004). [52] E. Hirakawa, H. Higuchi, and Y. Y. Toyoshima, Proc. Natl. Acad. Sci. USA 97, 2533 (2000). [53] C. Cho, S. L. Reck-Peterson, and R. D. Vale, J. Biol. Chem. 283, 25839 (2008). [54] J. L. Ross, H. Shuman, E. L. F. Holzbaur, and Y. E. Goldman, Biophys. J. 94, 3115 (2008). [55] S. J. King and T. A. Schroer, Nat. Cell. Biol. 2, 20 (2000). [56] G. Cappello, P. Pierobon, C. Symonds, L. Busoni, J. C. M. Gebhardt, M. Rief, and J. Prost, Proc. Natl. Acad. Sci. USA 104, 15328 (2007). [57] D. Tsygankov and M. E. Fisher, Proc. Natl. Acad. Sci. USA 104, 19321 (2007). [58] J. C. M. Gebhardt, A. E.-M. Clemen, J. Jaud, and M. Rief, Proc. Natl. Acad. Sci. USA 103, 8680 (2006). [59] A. E.-M. Clemen, M. Vilfan, J. Jaud, J. Zhang, M. Barmann, and M. Rief, Biophys. J. 88, 4402 (2005). [60] A. B. Kolomeisky and M. E. Fisher, Biophys. J. 84, 1642 (2003). [61] S. Uemura, H. Higuchi, A. O. Olivares, E. M. D. L. Cruz, and S. Ishiwata, Nat. Struct. Mol. Biol. 11, 877 (2004). [62] G. Oster and H. Wang, Nature 32, 459 (2000). [63] H. Yardimci, M. van Duffelen, Y. Mao, S. S. Rosenfeld, and P. R. Selvin, Proc. Natl. Acad. Sci. USA 105, 6016 (2008). [64] H. L. Sweeney and A. Houdusse, Curr. Opin. Cell. Biol. 19, 57 (2007). [65] Y. Oguchi, S. V. Mikhailenko, T. Ohki, A. O. Olivares, E. M. D. L. Cruz, and S. Ishi- wata, Proc. Natl. Acad. Sci. USA 105, 7714 (2008). [66] Z. Bryant, D. Altman, and J. A. Spudich, Proc. Natl. Acad. Sci. USA 104, 772 (2007). [67] M. Iwaki, A. H. Iwane, T. Shimokawa, R. Cooke, and T. Yanagida, Nat. Chem. Biol. 5, 403 (2009). [68] I. P. Udovichenko, D. Gibbs, and D. S. Williams, J. Cell Sci. 115, 445 (2002). [69] A. Inoue, J. Saito, R. Ikebe, and M. Ikebe, Nat. Cell. Biol. 4, 302 (2002). [70] M. Tominaga, H. Kojima, E. Yokota, H. Orii, R. Nakamori, E. Katayama, M. Anson, 26 T. Shimmen, and K. Oiwa, EMBO J. 22, 1263 (2003). [71] D.-E. Kim, M. Narayan, and S. S. Patel, J. Mol. Biol. 321, 807 (2002). [72] H. Risken, The Fokker-Planck Equation (Springer, Berlin, 1989). [73] Y. Zhang, J. Stat. Phys. 134, 669 (2009). [74] H. Wang and G. Oster, Europhys. Lett. 57, 134 (2002). [75] S. V. Gehlen, M. Evstigneev, and P. Reimann, Phys. Rev. E 77, 031136 (2008). [76] T. M. Nieuwenhuizen, S. Klumpp, and R. Lipowsky, Phys. A 350, 122 (2004). [77] A. B. Kolomeisky and M. E. Fisher, Ann. Rev. Phys. Chem. 58, 675 (2007). [78] S. Liepelt and R. Lipowsky, Phys. Rev. Lett. 98, 258102 (2007). [79] Y. Zhang, Physica A 383, 3465 (2009). [80] M. E. Fisher and A. B. Kolomeisky, Physica A 274, 241 (1999). [81] P. Reimann, C. V. den Broeck, H. Linke, P. Hanggi, J. M. Rubi, and A. P´erez-Madrid, Phys. Rev. Lett. 87, 010602 (2001). [82] Y. Zhang, Phys. Lett. A 373, 2629 (2009). [83] P. A. Pury and M. O. C´aceres, J. Phys. A: Math. Gen. 36, 2695 (2003). [84] A. B. Kolomeisky, E. B. Stukalin, and A. A. Popov, Phys. Rev. E 71, 031902 (2005). [85] P. Bieling, I. A. Telley, J. Piehler, and T. Surrey, EMBO Reports 19, 1121 (2008). [86] N. F. Endres, C. Yoshioka, R. A. Milligan, and R. D. Vale, Nature 439, 875 (2006). [87] R. Seidel, J. G. P. Bloom, C. Dekker, and M. D. Szczelkun, EMBO J. 27, 1388 (2008). [88] J. W. Shaevitz, S. M. Block, and M. J. Schnitzer, Biophys. J. 89, 2277 (2005). [89] Y. Q. Gao, Biophys. J. 90, 811 (2006). [90] M. Nishikawa, H. Takagi, T. Shibata, A. H. Iwane, and T. Yanagida, Phys. Rev. Lett. 101, 128103 (2008). [91] T. Masuda, BioSystems 95, 104 (2009). [92] E. Gerritsma and P. Gaspard, arXiv:0904.4218 (2009). [93] R. Lipowsky, Phys. Rev. Lett. 85, 4401 (2000). [94] R. Lipowsky and N. Jaster, J. Stat. Phys. 110, 1141 (2003). [95] Y. Chen, B. Yan, and R. Miura, Phys. Rev. E 60, 3771 (1999). [96] H. Y. Wang, C. S. Peskin, and T. C. Elston, J. theor. Biol. 221, 491 (2003). [97] H. Wang, Int. J. Numer. Anal. Model. 1, 1 (2004). 27 [98] R. D. Astumian, Science 276, 917 (1997). [99] A. Parmeggiani, F. Julicher, A. Ajdari, and J. Prost, Physical Review E 60, 2127 (1999). [100] J. M. R. Parrondo and B. J. D. Cisneros, Appl. Phys. A 75, 179 (2002). [101] P. Reimann, Phys. Rep. 361, 57 (2002). [102] M. Bier and R. D. Astumian, Phys. Rev. Lett. 71, 1649 (1993). [103] F. Julicher and J. Prost, Phys. Rev. Lett. 75, 2618 (1995). [104] J. Prost, J.-F. Chauwin, L. Peliti, and A. Ajdari, Phys. Rev. Lett. 72, 2652 (1994). [105] Y. Zhang, Chin. J. Chem. Phys. 23, 65 (2010). [106] B. Derrida, J. Stat. Phys. 31, 433 (1983). [107] H. Qian, Biophys. Chem. 67, 263 (1997). [108] The one-state model can be regarded as a simplification of the multi-state model but with an effective potential, Theoretically, this effective potential can be obtained by weighted average of potentials in each state of the multi-state model [97]. [109] For motor proteins, ωa(x), ωd(x) depend on the standard chemical potentials and concentrations of ATP, ADP and ionic phosphate Pi [2, 100]. 28 FIG. 1: Schematic depiction of two coupled one-dimensional hopping models. In which the forward and backward transition rates of molecular motor in state 1 are denoted by Fn and Bn, and are denoted by fn and bn for molecular motor in state 2, here 1 ≤ n ≤ N with N is the period of the hopping models. The inter-state transition rates at position n are denoted by ωn a (states 1→2) and ωn d (states 2→1). For motor proteins, ωn a , ωn d depend on the chemical potentials and concentrations of ATP and ADP. FIG. 2: Special case I of two coupled one-dimensional hopping models. In which ωn a = d = 0 for 1 ≤ n ≤ N − 1, and ωN ωn a = ωa, ωN d = ωd. For this special case, the mean velocity of molecular motor would be zero if there is no energy input in each of the two states, i.e., ∆G1 = ∆G2 = 0. In fact, at steady state, there is also no energy input during the process that drives the motor from one state to another, since ωaPN = ωdρN . 29 FIG. 3: Special case II of two coupled one-dimensional hopping models. In which ωn a = ωn d = 0 for n 6= M, N , and ωM a = ωa, ωM d = Ωd. For this special case, d = ωd, ωN a = Ωa, ωN the mean velocity of molecular motors might not be zero even if there is no energy input in each of the two states, i.e., ∆G1 = ∆G2 = 0. Since there usually exists energy input to molecular motor during its jump from one state to another unless ωaPM = ωdρM and ΩaPN = ΩdρN . FIG. 4: Different types of probability flux for special case II. From these figures, one can see that the motion of molecular motors is usually loosely coupled to the energy input process, i.e., there might exist energy input but without directed macroscopic mechanical motion. Part of the input energy will be consumed during substep oscillation.
1201.4966
1
1201
2012-01-24T13:03:48
Velocity distribution of neutral particles ejected from biological material under ultra short laser radiation
[ "physics.bio-ph", "physics.atm-clus", "physics.atom-ph", "q-bio.TO" ]
Neutral particles ejected from biological material under ultra short laser ablation have been investigated by laser post-ionization time-of-flight mass spectrometry. It could be shown, that beside ionized species, a substantial amount of neutral particles is ejected. A temporal study of the ablation plume is carried out by recording neutral particle time-of- flight mass spectra as a function of delay time between the ablation and post-ionization pulse. Close the ablation threshold, the mechanism of ejection is found to be of predominantely mechanical nature, driven by the relaxation of the laser-induced pressure. In this regime of stress confinement, the ejection results in very broad velocity distributions and extremely low velocities.
physics.bio-ph
physics
Velocity distribution of neutral particles ejected from biological material under ultra short laser radiation Wolfgang Husinsky∗ and Hatem Dachraoui† Institut fur Allgemeine Physik, Vienna University of Technology, Wiedner Hauptstrasse 8-10, A-1040 Wien, Austria Abstract Neutral particles ejected from biological material under ultra short laser ablation have been investigated by laser post-ionization time-of-flight mass spectrometry. It could be shown, that beside ionized species, a substantial amount of neutral particles is ejected. A temporal study of the ablation plume is carried out by recording neutral particle time-of-flight mass spectra as a function of delay time between the ablation and post-ionization pulse. Close the ablation threshold, the mechanism of ejection is found to be of predominantely mechanical nature, driven by the relaxation of the laser-induced pressure. In this regime of stress confinement, the ejection results in very broad velocity distributions and extremely low velocities. PACS numbers: 79.20.Ds, 79.20.Eb, 42.65.Re, 42.65.Sf. 2 1 0 2 n a J 4 2 ] h p - o i b . s c i s y h p [ 1 v 6 6 9 4 . 1 0 2 1 : v i X r a 1 I. INTRODUCTION The interaction of ultra short laser light with biological tissue is the basis for an im- mense and still expanding number of medical applications of lasers. Tissue processing with ultra short laser pulses has been of growing interest due to the high precision achieved. Refractive surgery inside the human eye or 'nano-surgery' of single biological cells are strik- ing examples1. The potential of ultra short laser radiation for surgical applications can be considered as established2,3. While the nature of the mechanisms behind ultrafast laser ablation of biological targets has been studied theoretically quite extensively, mostly using molecular-dynamics (MD) simulations4,5, the number of detailed experimental investigations is very limited. For ex- ample, molecular dynamics (MD) ablation studies of organic solids have suggested two principal classes of ablation mechanisms based on the laser parameters such as fluences and pulse widths: (i) photochemical processes in which the breakup of the material is the result of strong tensile stresses (spallation)6. (ii) photothermal ablation, where laser energy ab- sorption is followed by its conversion into heat, which finally results in high temperatures; possible consequences include homogeneous nucleation, phase explosion and vaporization7. For femtosecond laser pulse irradiation, the exact dynamics of the ablation process is not well understood as it appears to be a complex combination of these different mechanisms and also a function of the material properties which may change during the course of ablation. In this paper we analyze the mechanisms of material ejection from biological tissue under ultra short laser radiation, as well as their pulse width dependence. Using laser postion- ization mass spectrometry techniques we follow the temporal evolution of the plume com- position. The measured velocity distributions of the ejected neutral particles are extremely broad and cannot be described by a Maxwell Boltzmann distribution. For fluences close to the ablation threshold but below that for plasma formation, laser induced pressure due to overheating of the irradiated material is identified as the key processes that determine the dynamics of laser ablation. 2 II. EXPERIMENTAL SETUP The experiments have been performed in an UHV chamber equipped with a reflectron- type time-of-flight (TOF) mass spectrometer for detecting laser ablated particles. Laser- Post-Ionization is used for neutral particle detection. Ablation and post ionization is per- formed with ultra short laser radiation by a system which consists of two multipass CPA Ti: Sapphire amplifiers (Femtopower) seeded from a common mode locked Ti:Sapphire os- cillator (Femtosource). The system, operating at a repetition rate of 1 kHz, provides laser pulses at a center wavelength around 800 nm (1.5 eV photons) with a typical duration of 25 fs. The ablation beam was incident on the surface at an angle of 45 measured from the surface normal and was focused to a spot size of typically 100-500 µm in diameter. The ablation- and post-ionizing laser beam can be delayed with respect to each other (ablation and post-ionizing beam). The post-ionizing laser beam was crossing 2.5 mm normal in front of the target. The base pressure in the UHV chamber was approximately 10−7 mbar, since baking the chamber is not possible with biological samples. In addition, one of the two amplifiers was equipped with an opto-acoustic dazzler, which allowed pulse shaping of the laser beam. This measurement can be achieved despite the many experimental complexities that result from the not ideal vacuum conditions. Soft tissues (i.e. cornea), in particu- lar, pose problems in vacuum environments. We have chosen hard human tissues (bone, tooth) material for our investigations. The main components of human teeth are enamel and dentin. Enamel contains 95% hydroxyl-apatite (Ca10 (PO4)6 (OH)2), 4% water and 1% organic material. Furthermore, it contains impurities like Cl, Na, K, F or Mg. Dentin con- tains about 70% (Ca10 (PO4)6 (OH)2), 20% organic components (collagen fibers) and10% water. Human bone is made up of 50-60% hydroxyl-apatite, 15-20% water, 1% phosphates (inorganic components), 20% collagen and 2% proteins (organic components). III. RESULTS AND DISCUSSION Under ultra short laser irradiation, neutrals as well as ions are ejected from bone- and tooth-material. The detection and measurement of ablated ions is relatively simple and has been reported previously. The mass spectra of particles ablated from tooth material and bone material are very similar and have the basic features in common. Therefore it is sufficient 3 to display and to discuss the experimental results of one example, and we have chosen the data of tooth. A typical mass spectrum measured at laser fluence F = 200 mJ / cm2 is shown in Figure 1. The dominant signal (highest yield) arises for m/z = 40, with m ion mass and z the ion charge. In addition, the ion spectrum shows the presence of many weak peaks attributable to the presence of organic CmHn ions. The observation of relative large particles is a characteristic signature that photomechanical effects induced by pressure relaxation play a crucial role in the ablation of biological tissue. Further analysis was focused on the detection of the neutral particles. Typical mass spectra of the photoionized neutral particles obtained with fresh and etched ablated surfaces are shown in Figure 2. The spectra were measured for an ablation laser fluence of about 200 mJ/cm2 and at time delay ∆t = 5 µs between the ablation and the post ionization laser pulses. As can easily be seen, the mass spectrum is significantly modified. In addition to the large two peaks atm/z = 26 and m/z = 28, organics CmHn are observed. Only with the fresh surface, the organics CmHn are observed in relatively high abundance. In contrast to the cation spectrum the neutral mass spectrum shows no intensity at the mass/z (39, 96, 103 and 112). This observation suggests that many parent clusters undergo non linear dissociation during the post-ionisation event. Thus, the neutral spectrum consists essentially of fragment particles rather than direct ejected particles. Furthermore, comparison with the ion spectrum reveals, that close to the ablation threshold neutral and ions are ejected with a majority of neutrals. Given the large number of clusters found in the plume in this regime, ablation exhibits a mechanical rather than thermal character. Laser ablation in the short-pulse regime involves different processes which are effective at different deposited laser energies. In order to investigate the mechanisms of the ejection close to the ablation threshold we have measured the time-of-flight (TOF) distributions of several neutral particles. Contrary to measurements of the velocity of ejected ions, the measurement of neutrals is more informative due to the physical ablation process itself, since neutrals are not influenced by potentials in the surface (the surface-space-charge effect), which can obscure the results quite substantially and have led to different interpretations and velocities reported (eV-keV)8. The measurements are performed with etched surfaces when no signal change with the number of laser pulses occurred. Here it is important to note that the system relaxes within a few hundreds picoseconds after the excitation (<< the laser repetition period). Figure 3 shows the integration over the 28/z and 54/z peaks as a 4 function of the flight time between target and ionizing laser. The TOF spectra can be directly converted into a velocity or energy distribution. Both m/z = 28 and m/z = 54 distribution are maximized at 5 µs time delay that corresponds to 600 m / s velocity. This velocity of 600 m / s, observed for most particles, seem to be rather low at first sight, but similar slow velocity distributions have been predicted in earlier MD simulations of laser ablation of organic solids in the stress confinement irradiation regimes9. For all TOF data of biological samples measured, there is a clear discrepancy between the data and a best fit according to the Maxwell-Boltzmann (MB) distribution. The experimental TOF distributions are broader than a MB distribution. For comparison, we also include in figure 3 the TOF distribution of the emitted neutral Fe (m/z=56) from a steel target. In contrast to the biological TOF distributions, the measured Fe distribution can satisfactorily be described by a MB distribution. As we have clearly demonstrated previously in10 -- 12, in the case of a metal, matter removal can be attributed to phase explosion (photothermal effect). Based on these facts, photothermal effects are, under these conditions, not responsible for the ejection of material from biological tissue. For longer time delays ∆t > 20 µ s, neutral particles are observed in the mass spectra. Moreover within the experimental repetition period of 1 ms, the TOF distribution not recovers to the initial state. This feature was confirmed by observing neutral particles in the mass spectra recorded at negative delay time. Negative time delay corresponds to the post-ionization pulse preceding the ablation pulse. These extremely slow particles explain the observed background in the TOF distributions. This brings us to an obvious question: what is the origin of these particles? It is evident from the low temperature of the ejected particles that the physical processes leading to material ejection have predominantely mechanical character. Two observations support photomechanical mechanisms: (i) the neutral TOF distributions do not follow the MB distribution. (ii) the 28/z and 54/z peaks show the same temporal evolution; this constitutes a strong indication that the origin of the particles with low velocities (600 m / s) could be due to the ejection of larger clusters, which are fragmented by the postionization pulse into relatively small CmHn fragments. Our data is consistent with experimental13,14 and MD simulation9,15 results of laser ablation of biological tissue, where photomechanical effects has been identified as the main mechanisms responsible for the ablation in the regime of stress confinement. Furthermore, a plausible explanation of the observed neutrals particles 5 at time delay around 1 ms is the separation of a large surface layer from the sample with a velocity of 30 m/s. This behavior is very similar to that observed in simulations of ablation of organic solids16. In this theoretical modelling the authors demonstrated that irradiation under conditions of stress confinement can lead to the ejection of complete layers of the material. The origin of the broad distributions can be related primarily to layer (depth) effects. Confirmation can be found in the literature. Earlier molecular-dynamics simulations of ab- lation in organic solids have shown that different regions form in the sample during the ablation process. These regions differ in their expansion dynamics and in the pressure relax- ation they follow. This suggests that different ejection conditions for molecules, depending upon their original depth in the substrate. This means that the total TOF distribution might be a superposition of Maxwell-Boltzmann distributions with different stream veloci- ties. Moreover, subsequent collisions in the expanding plume can lead to further broadening of the distribution. The experimental TOF distribution (m/z = 28) can satisfactorily be described by a shifted Maxwell-Boltzmann distribution17. So, a fit of the TOF distribu- tions (m/z = 28) to a modified Maxwell-Boltzmann distribution results in temperatures of 1300 K. We can derive from the data a basic conclusion: the relevance of stress confinement as an important contribution to the ablation mechanism. In this scenario, the thermal expan- sion of the volume heated by a laser pulse generates compressive stresses. The subsequent propagation of these stresses from the free surface can transform them into tensile stresses of sufficient strength to cause mechanical fractures parallel to the surface of the sample and ejection of the upper layers. The spallation proceeds through nucleation, growth and coalescence of voids within the spallation region18,19. As already demonstrated in previous experimental and molecular dynamics simulation studies10,20, the ablation mechanisms, and the parameters of the ejected plume have a strong dependence on the laser pulse duration. Figure 4 shows the measured TOF distribution of the ablated neutral (m/z = 28) for two different additional laser pulse widths ∆τ = 350 and 500 fs (in addition to the data for the regular 25 fs pulses). The energy per pulse was kept constant. In this regime of stress confinement, pulse width variation allows us to obtain important information about the dependence of the material ejection on the laser peak intensity as a function of ∆τ . It is well known that the photomechanical fracture 6 is determined by the interplay between the tensile pressure and the thermal softening due to the laser heating: higher tensile pressure (up to -150 MPa) does not cause mechanical fracture19,21. Much broader TOF distributions with slow velocities are observed in Figure 4a, which can not be described by a Maxwell-Boltzmann distribution. For a pulse width ∆τ = 500 fs, the velocity distributions can be roughly described by a modified Maxwell-Boltzmann distri- bution. Simulations demonstrate that a broader distribution is characteristic for the plume ejected in the stress confinement regime9. Furthermore, a comparison with figure 3 (TOF distribution at 25 fs, circles ) reveals two interesting features in this measurement: (i) a clear shift of the TOF distributions to lower velocities. A similar effect of temperature-decrease has been observed earlier in MD simulations of laser ablation of molecular samples. The authors have related the decrease in the temperature of the plume to enhanced ejection and rapid material disintegration9, even though here a comparison between our results and the simulations has to be regarded with care (these MD simulations were performed for 15 and 150 ps duration pulses). We will return to this point below. (ii) Increase in the background value with increasing the pulse duration from 25 to 350 fs. This observation reflects the increase in the abundance of large cluster in the plume. The dependence of the amount of m/z = 28 material removed in the time window of about 1 ms versus laser pulse width is shown in figure 4b. The total ablated yield increases with increasing the laser pulse width. This observation can explain the observed decrease in the temperature of the plume (figure 4c). Our interpretation of the increase in the total amount of the ejected material takes into account two main effects: (1) An enhancement in the amplitude of the pressure wave and (2) under these radiation conditions, the balance between the amplitude of the tensile component and the thermal softening is more favourable for photomechanical effects (such as spallation). In the first scenario, in spite of the dominance of photomechanical ablation in the first case (short pulse ∆τ = 25 fs, high peak intensity), one can imagine that this regime is characterised by the occurrence of photothermal effects in addition to the photomechanical effects. Because laser-induced spallation can be initiated at energy densities much lower than those required for phase explosion and vaporization, we believe that the increase in the pulse duration leads to an irradiation regime below the thermal ablation threshold and thus phase explosion cannot occur and account for material ejection in this regime. So, 7 more laser energy is converted to mechanical energy of the thermoelastic stress wave. This is probably the reason why spallation becomes more dominant. Second, The etch depth is determined not only by the amplitude of the tensile component of the pressure wave, but also the temperature gradient produced within the irradiated surface region. Close to the ablation threshold a decrease of the peak intensity leads to an increase in the tensile pressure. This means that tensile-wave-mediated effects become more dominant for longer pulses (350 − 500 fs). IV. CONCLUSIONS In summary, we have investigated the mechanisms of the femtosecond laser ablation of biological material. At laser fluences above the ablation threshold, our results indicate that photomechanical effects, driven by the relaxation of high thermoelastic pressure, are re- sponsible for the collective material ejection. The velocity distributions are broader than a Boltzmann distribution and exhibit an offset (extremely slow particles), presumably frag- mentation products of larger clusters or surface layer. A comparison of the results the velocities obtained with 25 and 350 fs laser pulses also reveals a number of differences that can have important implications for practical applications of laser ablation. Larger ablated volume and broader velocity distributions are produced at the same laser fluences for pulse longer ∆τ = 350 fs than for τ = 25 fs. The experiments presented demonstrate the potential for further investigations to identify the ablation mechanisms of biological material at high laser fleunces. Experiments with soft tissues should follow. V. ACKNOWLEDGMENTS This work has been partly supported by the Austrian Science Foundation FW under project numbers P13756-N02 and P15937-N02. ∗ [email protected] 8 † Present address:Molecular and Surface Physics, Faculty of Physics, Bielefeld University, Ger- many 1 H. Lubatschowski, A. Heisterkamp, F. Will, A. I. Singh, J. Serbin, A. Ostendorf, O. Kermani, R. Heermann, H. Welling, and W. Ertmer, RIKEN Review 50, 113 (2002). 2 W. Sekundo and e. al., J Cat. Refract. Surg. 34, 1513 (2008). 3 H. Lubatschowski and A. Heisterkamp, in Femtosecond Technology for Technical and Medical Applications, Vol. 96 of Topics in Applied Physics, edited by H. L. F. Dausinger, F. Lichtner (Springer Verlag, Berlin, 2004), p. 91. 4 E. Leveugle, D. S. Ivanov, and L. V. Zhigilei, Applied Physics A 79, 1643 (2004). 5 E. Leveugle and L. V. Zhigilei, Applied Physics A: Materials Science &amp; Processing 79, 753 (2004). 6 P. Lorazo, L. J. Lewis, and M. Meunier, Physical Review B 73, 134108 (2006). 7 D. Perez and L. J. Lewis, Physical Review Letters 89, 255504 (2002). 8 O. Kreitschitz, W. Husinsky, G. Betz, and N. H. Tolk, Applied Physics A: Materials Science &amp; Processing 58, 563 (1994). 9 L. V. Zhigilei and B. J. Garrison, Journal of applied physics 88, 1281 (2000). 10 H. Dachraoui, W. Husinsky, and G. Betz, Applied Physics A: Materials Science &amp; Pro- cessing 83, 333 (2006). 11 H. Dachraoui and W. Husinsky, Applied Physics Letters 89, 4102 (2006). 12 S. Bashir, M. S. Rafique, and W. Husinsky, Applied Surface Science 255, 8372 (2009). 13 I. Itzkan, Proc.Natl Acad.Sci.USA 92, 1960 (1995). 14 D. Albagli, Opt.Lett. 19, 1684 (1994). 15 A. G. Zhidkov, L. V. Zhigilei, A. Sasaki, and T. Tajima, Applied Physics A: Materials Science &amp; Processing 73, 741 (2001). 16 L. V. Zhigilei and B. J. Garrison, Applied Physics A: Materials Science &amp; Processing 69, S75 (1999). 17 L. V. Zhigilei and B. J. Garrison, Applied Physics Letters 71, 551 (1997). 18 G. Paltauf and P. E. Dyer, Chemical Reviews 103, 487 (2003). 19 L. V. Zhigilei, E. Leveugle, B. J. Garrison, Y. G. Yingling, and M. I. Zeifman, Chemical Reviews 103, 321 (2003). 20 P. Lorazo, L. J. Lewis, and M. Meunier, Physical Review Letters 91, 225502 (2003). 9 21 D. Perez and L. J. Lewis, Physical Review B 67, 184102 (2003). 10 VI. FIGURECAPTIONS Fig.1: Typical mass spectrum of ions emitted from tooth sample under ultra short laser radiation (25 fs, 800nm F= 200 mJ/cm2). Fig. 2: Mass spectra of neutral particles of tooth sample produced at F= 200 mJ/cm2 and detected at ∆t = 5 µs with relatively fresh (a) and etched (b) surfaces. Fig. 3: Measured TOF distribution of m/z=28 and 54 neutral particles ablated from tooth material with 25 fs ablation laser pulse and fluence 200 mJ/cm2 (Triangles and circles). TOF distribution of ablated neutral Fe from steel for 400 fs laser pulse width and 100 mJ/cm2 laser fluence (squares). The red and green solid lines represent fits to a Maxwell-Boltzmann distribution. Blue line is a fit to shifted Maxwell-Boltzmann distribution. A 10-points moving average has been applied to the raw data. Error bars indicate the reproducibility of the data. Fig. 4: (A) Measured TOF distribution of m/z=28 for ablation laser pulse width τ = 350 and 500 fs and fluence 200 mJ/cm2. The lines represent fits to a modified-MB distribution. For comparison, we also include the modified-MB fit of the m/z=28 TOF distribution (blue line, figure 3)(B) Surface temperature, determined from modified MBD, as a function of the laser pulse width. (C) The total yields, derived from the area of the modified MBD, as a function of the laser pulse width. 11
1601.05260
2
1601
2017-07-16T17:27:50
On the perfomance of a photosystem II reaction centre-based photocell
[ "physics.bio-ph", "physics.chem-ph", "quant-ph" ]
The photosystem II reaction centre is the photosynthetic complex responsible for oxygen production on Earth. Its water splitting function is particularly favoured by the formation of a stable charge separated state via a pathway that starts at an accessory chlorophyll. Here we envision a photovoltaic device that places one of these complexes between electrodes and investigate how the mean current and its fluctuations depend on the microscopic interactions underlying charge separation in the pathway considered. Our results indicate that coupling to well resolved vibrational modes does not necessarily offer an advantage in terms of power output but can lead to photo-currents with suppressed noise levels characterizing a multi-step ordered transport process. Besides giving insight into the suitability of these complexes for molecular-scale photovoltaics, our work suggests a new possible biological function for the vibrational environment of photosynthetic reaction centres, namely, to reduce the intrinsic current noise for regulatory processes.
physics.bio-ph
physics
On the performance of a photosystem II reaction centre-based photocell Richard Stones, Hoda Hossein-Nejad, and Alexandra Olaya-Castro∗ Department of Physics and Astronomy, University College London, Gower Street, London, WC1E Rienk van Grondelle Department of Physics and Astronomy, VU University, 1081 HV Amsterdam, The Netherlands (Dated: September 24, 2018) The photosystem II reaction centre is the photosynthetic complex responsible for oxygen pro- duction on Earth. Its water splitting function is particularly favoured by the formation of a stable charge separated state via a pathway that starts at an accessory chlorophyll. Here we envision a photovoltaic device that places one of these complexes between electrodes and investigate how the mean current and its fluctuations depend on the microscopic interactions underlying charge separa- tion in the pathway considered. Our results indicate that coupling to well resolved vibrational modes does not necessarily offer an advantage in terms of power output but can lead to photo-currents with suppressed noise levels characterizing a multi-step ordered transport process. Besides giving insight into the suitability of these complexes for molecular-scale photovoltaics, our work suggests a new possible biological function for the vibrational environment of photosynthetic reaction centres, namely, to reduce the intrinsic current noise for regulatory processes. Life on Earth is fueled by photosynthesis, the process by which plants, algae and certain bacteria convert solar energy into stable chemical energy1. The initial electron transfer steps during solar energy conversion by these organisms take place in photosynthetic reaction centres (PRCs), sophisticated trans-membrane supramolecular pigment-protein complexes that exhibit a dual device-like functionality. Under illumination, a PRC complex effec- tively operates as Nature's solar cell1 where electronic excitations of chromophores are transformed into stable charge-separated states, with electron donor and elec- tron acceptor separated by a few nanometres. This pi- cosecond charge separation process occurs with near unit quantum efficiency implying that almost every quanta of energy absorbed results in charge separated across the PRC2,3. The same chromophore-protein structure and energetic landscape promoting this quantum yield also favours a diode-like behaviour of all PRCs such that un- der an appropriate applied bias, electric current flows almost entirely in one direction4. Their functional versa- tility, nanometre size, and near-unit quantum efficiency has motivated the exploration of PRCs as possible com- ponents of photovoltaic and photoelectrochemical cells5,6 as well as in biomolecular electronics4,7,8. A step further in this field is the recent development of single-molecule techniques that allow measurement of the photocurrent through individual PRC complexes9. Using cysteine group mutations, it has been possible to bind a photosystem I unit to a gold substrate and use a scanning probe gold-tip that acts as both an electrode and a localized light source to excite and measure the photocurrent of a fully functional PRC9. Moreover, us- ing a tapping mode atomic force microscope, it has been possible to confirm that electrons tunneling through a bacterial PRC, under an applied bias, follow the transfer pathway of the photochemical charge separation4. These experiments open up a new platform to carry out fur- ther investigations on how the microscopic mechanisms underlying the function of PRCs affect the electric cur- rent output of a single PRC, as well as to reveal further details of such microscopic mechanisms. In particular, it is foreseeable that besides measuring the current-voltage features these techniques may allow characterization of current fluctuations and the associated counting statis- tics of electron transport in PRCs. In quantum trans- port setups10 -- 12 it has been shown that such fluctuations can reveal intrinsic dynamical features of the quantum system through which electron transport occurs, includ- ing the influence of electron-phonon interactions13 and coherence14. A theoretical study along the lines of count- ing statistics for light-harvesting complexes has been car- ried out15 but so far there has been no investigation of full counting statistics of charge transport in PRCs. The photosystem II reaction centre (PSIIRC), present in higher plants, algae and cyanobacteria1, is arguably the most important prototype to be considered as it is responsible for water splitting and production of all oxy- gen on Earth16. Experimental evidence indicates that one of the distinguishing features of PSIIRC is the ex- istence of at least two different charge separation path- ways, one of which starts in the monomeric chlorophyll of the active D1 branch (ChlD1)17 -- 19. This transfer path- way has been argued to be a deciding factor for the func- tional operation of PSIIRC as a water splitting complex20 and is therefore the focus of this paper. While the de- tailed quantum mechanical features underpinning charge separation in PRCs are still under scrutiny20, there is a wealth of steady-state and time-resolved spectroscopy revealing the electronic state space and spectral density of fluctuations relevant for the formation of stable charge separated states19 -- 22. However, the implications of these microscopic mechanisms for the electric current output of a single PSIIRC are largely unknown. In fact, it is unclear whether these natural light-to-charge converting units are well suited for anthropogenic use: would pho- tocells integrating the microscopic mechanisms of PRCs deliver optimal power? How do these mechanisms affect the statistics of electron transport? The answers to these questions hold the potential to provide valuable insight both on the biophysics of these systems and for the de- velopment of the next generation of bio-inspired energy technologies. In this work we address these questions by envisioning a photocell device where a single PSIIRC using the ChlD1 pathway is placed between two electrodes and investigate how the microscopic mechanisms underlying the photo- chemical charge separation affect the electric current out- put and its fluctuations, under continuous incoherent il- lumination. By comparing the photocell operation under different spectral densities characterizing the interaction between electronic and vibrational degrees of freedom, we show that selective coupling to underdamped vibra- tions does not necessarily offer an advantage in terms of current and power output for this photocell but they lead to output currents with suppressed noise levels as quantified by a Fano factor less than one. A structured spectral density that includes coupling to well-resolved vibrations allows the noise strength to probe the struc- ture of the exciton manifold which transfers population to charge transfer states and leads to a sub-Poissonian statistics. This indicates that both the exciton mani- fold and the electron-vibration interactions in PSIIRC support a multi-step ordered electron transport process. Our work therefore puts forward a new possible func- tional role for the vibrational environment of PRCs, that of guaranteeing intrinsic current noise control for regu- latory purposes. Our analysis also gives insight into the suitability of PSIIRCs and their microscopic principles for photovoltaic or nano-electronic applications. RESULTS A photocell based on the photosystem II reaction centre The prototype complex we consider is the PSIIRC for which crystallography has provided the arrangement of the chromophores involved in primary charge transfer23. As illustrated in Fig. 1 (a), four chlorophylls (Chl) and two pheophytins (Phe) are arranged in two branches (D1 and D2), where D1 and D2 label the chlorophyll bind- ing proteins in the core of the reaction centre. The two central chlorophylls PD1 and PD2 (known as the special pair) are flanked by the accessory chlorophylls ChlD1 and ChlD2, and the pheophytins PheD1 and PheD2. Charge separation only occurs down the D1 branch24,25. Nonlin- ear spectroscopy has revealed at least two different ex- cited states, (PD1PD2ChlD1)∗ and (ChlD1PheD1)∗, that give rise to two different pathways (denoted the PD1 and ChlD1 pathways respectively) for charge separation along the D1 branch19,21,26. The likelihood of each depends on the specific protein configuration and the corresponding disorder of pigment excitation energies19. 2 Although the relative contribution of these pathways in ensemble measurements is not conclusively known, spec- troscopy and its corresponding theoretical fit indicate that electron transfer is predominantly initiated from the state (ChlD1PheD1)∗26. This transfer pathway is con- sidered an important building block for the functional operation of PSIIRC because by placing the ChlD1 as the lowest energy pigment it encourages charge trans- fer through D1 and by lifting the energy of the PD1PD2 pair it provides favorable conditions for the water oxi- dation function20,27. Our analysis therefore focuses on this ChlD1 pathway in which inter-pigment electronic couplings lead to formation of delocalised exciton states upon photo-excitation as revealed by spectroscopy19,28. However, for this protein configuration there is no ev- idence of coherent coupling between excited states and charge transfer (CT) states19,21,26, unlike the case of the PD1 pathway where ultrafast non-linear spectroscopy has given evidence of coherent electron transfer19,29. This is consistent with a very weak electronic coupling be- tween the low energy ChlD1 and PheD1 pigments and the primary CT state (see Table S2). The ChlD1 path- way can thus be seen as the relaxation dynamics in the energy landscape illustrated in Fig. 1(b). Population of an exciton state with the largest amplitude in the pair (ChlD1PheD1)∗ undergoes a dynamics within the exci- ton manifold while it is incoherently channeled towards an initial CT state Chl+ D1Phe− D1i and from there to the stable secondary CT state P+ D1Phe− D1i in which electron and hole reside on different chromophores separated by a few nanometres along the D1 branch (see Fig. 1(a)). Fluorescence line narrowing experiments30 have given evidence of a highly structured spectral density charac- terizing the interactions of an excited chromophore in PSIIRC with a wide range of vibrational motions as shown in Fig. 2. Although some of the sharp modes in this spectral density have frequencies that match en- ergy gaps between exciton states29, we will show that this feature is not entirely determinant for the mean current output of a PSIIRC-based photocell operating within the ChlD1 pathway. Together with the state-space aforemen- tioned, this spectral density of fluctuations has provided a good fit for steady-state and transient spectroscopy of primary charge separation in PSIIRC21,26,29. We therefore consider these features of the ChlD1 charge separation pathway to put forward two key ques- tions: (i) do these naturally occurring electron-vibration interactions and their associated spectral density provide the best strategy for maximizing current output in a bio- inspired photocell? and (ii) how do the exciton manifold and the vibrational environment affect the statistics of electrons flowing through this PRC? These questions are addressed by envisioning a theo- retical photocell device in which a single PSIIRC unit operating with the ChlD1 pathway is placed between two leads that can supply or take away electrons from the system, as illustrated in Fig. 1 (a). The charge transfer cycle is such that electrons are pumped from the source 3 (a) (b) ~62meV ~34meV ~50meV 1.4eV FIG. 1. Photosystem II reaction centre photocell model. (a) Schematic diagram of a proposed experimental setup for the photocell unit. The isolated core chromophores of PSIIRC are positioned between a gold substrate and a gold coated scanning probe microscope tip which act as electrodes. A silicon field-effect transistor (SFET) placed near the drain electrode could be used to measure the current statistics. (b) Energy level diagram showing the electronic state space of the model. The red arrows represent rates connecting the ground and empty state to the excited state manifold. ΓL and ΓR connect the system to the source and drain leads respectively while γex represents a coupling to an optical field which excites the system from the ground state to the lowest energy exciton state. Green arrows represent Forster/Marcus rates for primary and secondary charge separation. The load between states α and β indicates the transition across which we calculate the output current of the photocell and its statistics. s t i n u . b r a 1 0 0 s t i n u . b r a 12 10 8 6 4 2 0 0 J1 (ω) J2 (ω) 200 400 600 800 1000 frequency (cm−1 ) JM (ω) G1 (ω) G2 (ω) 50JD (ω) 400 800 1200 1600 frequency (cm−1 ) FIG. 2. Photosystem II reaction centre spectral den- sities. The components of the spectral densities used in the PSIIRC photocell model. Mode parameters of the structured component JM (ω) are shown in Table S1. The Drude part JD(ω) is scaled relative to the high energy parts for clarity. The inset shows the spectral densities J1(ω) and J2(ω) which are used to approximate the full spectral density. lead at rate ΓL and leave the system from the drain lead at rate ΓR while the sample is incoherently photo- excited at rate γex. Our model enforces the Coulomb blockade regime such that the probability of two elec- tron occupancy of the photocell is negligible31. We also consider an infinite applied bias that assures unidirec- D1Phe− D1Phe− tional electron flow32,33. Assuming low enough excita- tion rates to guarantee that only single excitation states are populated, the state space of the PSIIRC-based pho- tocell, shown in Fig. 1 (b), spans the following: the ground state gi, six exciton states X1i to X6i, the ini- tial CT state Chl+ D1i ≡ Ii, the secondary CT state P+ D1i ≡ αi and the positively charged state P+ D1PheD1i ≡ βi which represents the 'empty' state of the system for counting statistics calculations. The six exciton states arise from diagonalization of the site part of Hamiltonian Hel (see Methods) which includes coher- ent electronic interactions among all the six core chro- mophores located in both the D1 and D2 branches of the PSIIRC. Although the D2 branch is not directly involved in charge separation, excitons localized here can act as electronic traps26 and therefore can affect the statistics of electrons flowing through the system as we will discuss later. To investigate the effects of the vibrational environ- ment in the performance of the photocell, we aim to com- pare four cases corresponding to the four different spec- tral densities depicted in Fig. 2: (i) the case where the full structured spectral density J(ω) = JD(ω) + JM (ω) is considered, the cases where, besides the low-energy back- ground JD(ω), we account for one and two well-resolved modes with spectral densities (ii) J1(ω) = JD(ω)+G1(ω) and (iii) J2(ω) = J1(ω) + G2(ω), and (iv) the case where only the smooth low energy component JD(ω) is in- cluded. The expressions for the different components are given by21: JD(ω) = 2λDωΩD ω2 + Ω2 D , Gj(ω) = 2λjω2 j γjω j ω2 . (ω2 − ω2 j )2 + γ2 (1) (2) JD(ω) is the Drude form of a spectral density describing an overdamped Brownian oscillator where λD and ΩD are the reorganisation energy and cut off frequency, re- spectively. Gi(ω) describes the spectral density of an un- derdamped mode coupled to an excited pigment, with λj, ωj and γj being the reorganisation energy, frequency and damping rate of mode j respectively. JM (ω) = Pj Gj(ω) has been measured experimentally30 and includes 48 un- derdamped modes. For case (ii) described by J1(ω), we consider ω1 = 342cm−1 and for case (iii) correspond- ing to J2(ω), we consider ω1 as well as ω2 = 742cm−1. These two modes have been argued to be important for electron transfer along the alternative PD1 charge sepa- ration pathway29 and in our case their frequencies span all the energy gaps of the system. Hence, J2(ω) pro- vides a good approximation to the full spectral density. All parameters for these spectral densities are detailed in Supplementary Note 1. By comparing these cases we at- tempt to address the question of how well 'adapted' are these electron-vibrational interactions for photovoltaics: if it were possible to decouple these well resolved nuclear motions from the charge separation process, would the resultant photocell exhibit a better current and power output? Besides its theoretical relevance, this compari- son is experimentally motivated as such decoupling may be feasible via optical cavities34,35. As we will discuss in the next section, the mean current and power of our photocell is determined by the steady- state population of the secondary CT state αi. A non- perturbative computation of the steady-steady state un- der the influence of the full spectral density J(ω) with its 48 sharp modes per electronic state is quite challenging and out of the scope of our computational capabilities. We can however compute non-perturbative dynamics and steady state of our photocell including coherent interac- tions among single-excitation states and under the influ- ence of JD(ω), J1(ω) and J2(ω) using the hierarchical equations of motion36 -- 38 (see Methods). For compari- son we also investigate the dynamics and steady state predicted by a simpler Pauli master equation for state populations, with transfer rates as described in Ref.26 (see Methods). Theoretical justification of the validity of this approximate framework is presented in Supplemen- tary Note 3. Figures S2 and S4 show that the popula- tion dynamics and steady-state populations predicted by the accurate framework and the Pauli master equation agree qualitative and quantitatively. Furthermore it has been shown that this approximate scheme accurately re- produces transient and steady-state spectroscopy of the PSIIRC26. The main difference we see is that, as ex- pected, the accurate framework predicts some short-lived 4 excitonic coherences (see Fig. S3) which we show to have negligible influence on the current and power delivered by our photocell. Photocell current and power performance We fix the rate ΓL at which electrons are injected and set γex to simulate excitation by concentrated so- lar radiation39,40, ensuring detailed balance as specified in Supplementary Note 2. The generated steady-state current passing between the system and drain electrode is equivalent to the current flowing across a hypothetical load connecting the final states αi and βi, which have an associated energy gap Eαβ = Eα − Eβ. The voltage V across such a load quantifies the extractable energy from our photocell with final energy gap Eαβ and an expres- sion for V can be derived following standard thermody- namic considerations of photocells41 and photochemical systems42. Denoting the steady state populations of the secondary CT state ραα and the 'empty' state ρββ, the load voltage V can be expressed as: eV = Eαβ + kBT ln(cid:20) ραα ρββ(cid:21) , (3) where kB is the Boltzmann constant, T the temperature of the photocell and e the electric charge. The aver- age current hIi and the power output P delivered by the photocell are given, respectively, by hIi = eΓRραα and P = hIiV . By fixing all parameters except the rate ΓR at which electrons leave the system, we can then investigate the characteristic hIi − V and P − V curves which define the photovoltaic performance of a photocell. Figure 3 presents the characteristic curves for the four spectral densities depicted in Fig. 2. For JD(ω), J1(ω) and J2(ω), we present the results obtained both by the hybrid framework and by the approximate Pauli mas- ter equation, showing their remarkable agreement for the current and power predictions (consistent with the dy- namics shown in Figs. S2 and S4). In all the cases, the limit of ΓR → 0 leads to hIi → 0 defining the maxi- mum available voltage or open circuit regime eVoc which is proportional to the energy gap between the ground state and the state that is directly photo-excited40 i.e. eVoc ≈ E1 − Eg ≈ 1.8 eV . In the opposite limit, when ΓR → ∞, V → 0. In these two extremes the photocell delivers no power. For all spectral densities we observe that the current is constant at low voltages and drops off at a characteristic voltage V comparable to Eαβ when the spectral density is JD(ω). This characteristic voltage increases slightly as the spectral density includes more well-resolved modes. There are two remarkable features to highlight in Fig. 3. First, the constant current observed for small volt- ages and the maximum power are significantly lower for a photocell with a structured environment. We observe such a reduction even in the case of J1(ω) whose central frequency is quasi-resonant with several exciton energy gaps. This contrasts with the power enhancement pre- dicted for a simple light-harvesting unit operating under coherent interactions between all states43. Second, the current delivered by a photocell with J2(ω) is already quite close to that of J(ω), confirming that for the per- formance of the photocell, J2(ω) is a good approximation to the full spectral density. The behaviour predicted for the current and power de- livered by the PSIIRC photocell relies on two main fea- tures characterizing the ChlD1 charge separation path- way: the large reorganisation energies of the CT states and the concomitant weak coupling between the primary CT state and single-excitation states. These two facts lead to an incoherent population dynamics that is well described by second order perturbation theory in the electronic coupling and such that the rates of transfer between an exciton Xi and the primary CT state Ii are dominated by the convolution of their line shape functions44,45. Specifically, the rate of transfer between these states is given by kX,I = VX,I 2SX,I where VXI is the effective electronic coupling between the states and SXI quantifies the spectral overlap: SXI = 2ReZ ∞ 0 dteiωXI te−i(λX +λI )te−(gX (t)+gI (t)), (4) with ωXI the energy gap between the states, λX(I) the corresponding reorganisation energies for each state and gX(I)(t) the associated line broadening functions. Full expressions for these functions can be found in Supple- mentary Note 2. Figure S5 shows the overlap between the low-lying exciton X1i and Ii for the three spectral densities. As the spectral density contains more peaks, the reorganization energy of both states increases. How- ever, for all the cases the values for λI are about one order of magnitude larger than λX (see Table S5). The larger λI , the wider is the shift between donor and acceptor states and as such the overlap SXI accounting for spec- tral resonances is reduced to yield a lower rate of transfer kX,I . Similar considerations apply for the transfer rate between the CT states Ii and αi whose reorganisation energies satisfy λI < λα as discussed in Supplementary Note 4. In this scenario the condition of quasi-resonance between electronic gaps and well-resolved vibrations be- comes irrelevant for population transfer. The reduced transfer rates from excitons to the intermediate CT state Ii can then be interpreted as a Zeno-like effect whereby the strongly coupled environment "measures" the popu- lation of the CT state at a very high rate thereby slow- ing transfer. The disparity between reorganisation ener- gies for CT states and the donor exciton states ensures a downhill relaxation that in the biological context coun- teracts charge recombination. Zero-frequency noise 5 JD (ω) (hybrid) J1 (ω) (hybrid) J2 (ω) (hybrid) JD (ω) (Pauli) J1 (ω) (Pauli) J2 (ω) (Pauli) J(ω) (Pauli) 1.3 1.4 1.5 1.6 1.7 1.8 voltage (V) JD (ω) (hybrid) J1 (ω) (hybrid) J2 (ω) (hybrid) JD (ω) (Pauli) J1 (ω) (Pauli) J2 (ω) (Pauli) J(ω) (Pauli) (a) 2.0 1.5 e \ t n e r r u c 1.0 0.5 0.0 0 (b) 3.0 e / r e w o p 2.5 2.0 1.5 1.0 0.5 0.0 0.0 0.2 0.4 0.6 0.8 1.0 voltage (V) 1.2 1.4 1.6 1.8 FIG. 3. Photocell mean current and power versus volt- (b) age for different spectral densities. Power. Solid lines are calculated using HEOM hybrid model while dotted lines are calculated using the Pauli master equa- tion. Calculations carried out at 300K with excitation rate γex = 75cm−1. See Supplementary Note 1 for all other pa- rameters. (a) Current. output. Steady-state current fluctuations quantified by second and higher order cumulants (hI nic with n > 1) can give information on the microscopic mechanisms un- derlying correlations between elementary charge trans- fer events10. While the theory of full counting statistics is well established for Markovian systems31,32, for non- Markovian dynamics only a few comprehensive frame- works based on perturbative approaches have been put forward46. To address this shortcoming, we have devel- oped a non-perturbative formalism to compute the full counting statistics47, which integrates the exact system dynamics provided by the hierarchical equations of mo- tions with a recursive scheme that allows accurate com- putation of the current cumulants46 (see Methods). We focus on the long-time limit or zero-frequency regime of the relative noise strength which is quantified by the second order Fano factor48: Full performance characterization of a photocell can- not be limited to the steady-state current and power F (2) = hI 2ic hIi . (5) This ratio between the zero-frequency second order cur- rent cumulant and the mean current quantifies the devi- ation of the underlying statistical process from a Poisso- nian distribution. A Fano factor of 1 indicates a Poisso- nian process without correlations among charge transfer events. Deviations from 1 are interpreted as either super- Poissonian (F (2) > 1) or sub-Poissonian (F (2) < 1), regimes that can be associated with highly fluctuating or more stable currents, respectively. Figure 4 reports both non-perturbative and approxi- mate results for F (2) versus V at room temperature for JD(ω) (Fig. 4(c)) and J2(ω) (Fig. 4(b)). For simplic- ity, we have omitted the results for J1(ω) as they simply follow the same trend. The qualitative and quantita- tive agreement between these curves indicate that zero- frequency noise properties of the photocell for these spec- tral densities is dominated by a population dynamics that is well-captured by the approximate Pauli frame- work. This is also consistent with the fact that exciton coherences arising from the interaction with a slowly re- laxing bath or well-resolved vibrational motions decay much faster than the time scale on which the steady state is reached as discussed in the supplementary information. These arguments then justify the use of the approximate scheme to obtain insights into the behaviour of F (2) for the full spectral density J(ω) as shown in Fig. 4(b). the In mesoscopic and quantum systems zero- frequency Fano factor has proven to be very sensitive to the structure of the state space49 as well as to system- environment interactions13. This is precisely what is in- dicated by the results for the Fano factor of the cur- rent through our photocell device shown in Figs. 4 (a) and (b), which show that the noise profiles for J2(ω) and J(ω) each have a single minimum but are not symmet- ric with respect to the voltage at which this minimum occurs. In both cases we see that for V → 0 we have F (2) < 1 while in the opposite limit of V ≫ Eα,β we obtain F (2) → 1. This indicates that at small voltages the electron transport is slightly correlated in all cases. As the spectral density exhibits more structure, the noise levels are overall lower. For instance, for the full spec- tral density we have F (2)(V = 0) = 0.90 while for J2(ω), we have F (2)(V = 0) = 0.95. Similarly, the minimum of F (2) reaches lower values as the spectral density ac- quires more structure i.e. F (2) = 0.55 for J(ω) which is less than the values obtained for J2(ω) i.e. F (2) = 0.6 and for JD(ω) i.e. F (2) = 0.7. In all the cases the sub- Poissonian behaviour is a manifestation of the Coulomb blockade regime where the presence of an electron in the system prevents another one entering until the sys- tem is empty. However, the more "ordered" transport observed for spectral densities with more well-resolved spectral features relies on the rapid population transport among excitons induced by such structured vibrational environments. The rates of transport to CT states, while reduced, are still comparable to the transfer among ex- citons such that the statistics of transitions from states αi to βi samples the manifold of exciton states donating 6 population to the primary CT state. This hypothesis is confirmed by our analysis of the energy scale determining the voltage at which F (2) is minimum in each case of Fig. 4. The analytic form of the Fano factor for our photo- cell model is too cumbersome to give any insight into the conditions determining the minima in Fig. 4. In order to rationalise the minimum in each curve it is use- ful to consider the case of a single resonant level (SRL) in the infinite bias limit32. The dynamics of this sys- tem in the basis {occupiedi, emptyi} is governed by a Liouvillian with matrix elements L11 = −L21 = ΓR and L22 = −L12 = ΓL such that the Fano factor as a function of the voltage of a load across the occupied and empty states (Eq. (3)) takes the form F (2)(V ) = 1 + exp[−2(eV − E0)/kBT ] (1 + exp[−(eV − E0)/kBT ])2 , (6) L+Γ2 R (ΓL+ΓR)2 (cf. Eq. where E0 is the energy gap between the occupied and empty states. It is simple to show that this expres- sion is equivalent to writing the Fano factor in terms of ΓL and ΓR as F (2) = Γ2 (45) in Ref.32 and see Supplementary Note 6). Figure S2 shows that F (2)(V ) for the SRL exhibits a single minimum, just as in our PSIIRC photocell. The minimum occurs when Vmin = E0 which is equivalent to the condition of ΓL/ΓR = ρoccupied/ρempty = 1 as shown in Sup- plementary Note 6. In this case, however, the function is symmetric about Vmin approaching 1 at both large and small voltages and indicating that electron transfer events in these extremes are uncorrelated. Based on this, we can say that near the Vmin the noise in our PSIIRC is approximately equal to that of an effective SRL with occupied level α∗i, empty level βi and renormalized en- ergy gap E∗ αβ that determines Vmin. Denoting as Ejk the energy gap between states ji and ki of our photo- cell, we notice that for the case of JD(ω) the Fano factor has a minimum for Vmin ≈ EIβ ≈ 1.50eV while for the full spectral density Vmin ≈ EX6β ≈ 1.56eV and for the two mode spectral density J2(ω) we have, as expected, a value in between. This indicates that in the photocell with the full spectral density, the minimum noise samples the largest energy gap between βi and the exciton man- ifold while for JD(ω) the noise only witnesses the energy gap up to the intermediate CT state. This is consistent with the fact that the rate of transfer among excitons in the D1 branch are larger for J(ω) than for the other two spectral densities. As mentioned above, the most important feature of Figs. 4 (a) and (b) is the non-symmetric profile F (2) with respect to Vmin. We now show that the rate limiting this asymmetric behaviour is the secondary charge transfer rate. To do this we consider the situation where the rate of secondary CT transfer kI,α is set by hand to a very low value compared to relaxation rates within the exciton manifold and the rates between the excitons and primary CT state for the full spectral density. In this (a) (b) (c) ) 0 ( ) 2 ( F ) 0 ( ) 2 ( F ) 0 ( ) 2 ( F 1.0 0.9 0.8 0.7 0.6 0.5 1.0 0.9 0.8 0.7 0.6 0.5 1.0 0.9 0.8 0.7 0.6 0.5 0 J(ω) J2 (ω) JD (ω) Pauli hybrid 1.3 1.4 1.5 1.6 1.7 1.8 voltage (V) FIG. 4. Fano factor versus voltage. (a) Fano factor for the PSIIRC photocell with the structured spectral density J(ω) for a modified (slow) (dotted line) and the measured (solid line) secondary charge transfer rate. (b) Fano factor for the PSIIRC photocell with spectral density J2(ω) containing two underdamped vibrations. (c) Fano factor for the PSIIRC photocell with a smooth low-energy vibrational environment JD(ω). Calculations carried out at 300K with excitation rate γex = 75cm−1. See Supplementary Note 1 for all other pa- rameters. case, the population of αi is so slow that all the internal transfers from the exciton manifold to the primary CT can be described as a single step process i.e. there is no sampling of the exciton manifold and the Fano factor tends to 1 for small and larger voltages as shown by the dotted line in Fig. 4 (a). This means that the system behaves as a SRL for all values of V with renormalized Γ∗ L (cf. Fig. 4 (a) with Fig. S6). The time scale of secondary charge transfer is therefore a limiting time- scale which can lead to a variety of phenomena as will be further explored in the next section. To conclude, our results at room temperature indicate that the PSIIRC-based photocell with the structured vi- brational environment delivers less power than a pho- tocell with an unstructured environment, yet this is ac- companied by a suppression of current fluctuations. This noise reduction is a consequence of the different features of vibration-assisted transport among excitons and trans- port between excitons and charge transfer states, which underlies the function of our model PSIIRC. DISCUSSION Present single-molecule technologies demonstrating the ability to manipulate and measure the photocurrent of single PRCs have motivated us to investigate how the microscopic physical mechanisms underlying the function of these complexes may affect their performance as com- ponents in a photovoltaic cell. To approach this ques- 7 tion we have brought together biological and physical perspectives by considering a PSIIRC-based photocell in a protein configuration that is argued to favour water- splitting20 and in which charge separation is initiated at 17 -- 19. This protein configuration is the accessory ChlD1 characterized by the lack of coherent delocalization be- tween excitons and CT states19. Our results show that a structured environment assist- ing electron transfer in our model PSIIRC-based photo- cell device acts to reduce the current and power output in comparison to a situation where electron-vibration in- teractions are described by a simple smooth low-energy background function. This reduction is a manifestation of a Zeno-like effect whereby the weak donor-acceptor electronic coupling concomitant with the stronger cou- pling of CT states to well-resolved high energy modes lead to slower transfer rates to CT states. These obser- vations suggest that while PSIIRC complexes operating in the ChlD1 pathway may favour water oxidation un- der in vivo conditions, they may not necessarily be well suited for maximizing current output in single-molecule photovoltaics. Notwithstanding, the predicted reduction in the average photocurrent upon inclusion of coupling to well-resolved high-energy modes is not detrimental for the biological operation of PRCs. In the biological sce- nario it is more important to inhibit charge recombina- tion and to ensure the captured energy is not wasted. In the ChlD1 transfer pathway, stronger coupling of CT states to these well-resolved vibrational motions ensures downhill relaxation thereby helping to prevent charge re- combination. For anthropogenic purposes of obtaining the largest current out of these units regardless of its fluctuations, the best strategy then may be to decouple specific vibra- tional motions from electronic states. Indeed, modifica- tions of the electron-nuclei interactions can be achieved by strong coupling of pigment-protein complex to a con- fined optical cavity mode34,35. In particular, Ref.35 has shown that the energy exchange of electronic transitions with a strongly coupled optical mode could help suppress- ing reorganisation energy of the nuclei thereby increas- ing the rate of electron transfer reactions. Alternatively, one can select PSIIRCs operating in the PD1PD2 path- way where coherent delocalization across exciton and CT states has been probed19,29,50,51 and which will lead to an enhancement as predicted in Ref.43. While no advantage is obtained in terms of mean cur- rent and power output, strong coupling to well resolved vibrational modes results in a reduction of current fluc- tuations of our PSIIRC-based photocell. This lower noise strength obeying a sub-Poissonian statistics and the as- sociated ordered electron transport is promoted by the exciton manifold and signals out the multi-step nature of the transport process. Preliminary calculations (not shown) for a photocell with delocalized states across ex- citons and CT states indicate that such a noise reduction maybe a general feature. From the electronic-device viewpoint, reducing any kind of noise is always a desirable feature to guaran- tee device resolution; this includes intrinsic noise due to the inherently probabilistic nature of the process. Hence the device functionality of this noise reduction appears to be straightforward: to improve precision in the cur- rent delivered. More interesting is to discuss the possible advantages of such noise reduction in the biological con- text. It is well known that noise and its control is crucial across all scales in biology52,53. For instance, it has been discussed that biochemical processes that are inherently stochastic include mechanisms to control intrinsic noise and, in particular, to reduce it for regulatory processes53. Indeed, electron transfer events in photosynthetic reac- tion centres belong to a larger family of stochastic trans- port processes in biology, some of which have already been predicted to exhibit mechanisms suppressing fluc- tuations below the Poisson level54. Moreover, increased complexity in biological networks has been linked to in- trinsic noise reduction55. We therefore argue that, for bi- ological function of PRCs, the coupling to well-resolved vibrations and the predicted noise reduction could in- deed have a regulatory function. In these systems, the final stable CT state αi donates an electron to quinone B and once this reduction happens, the PRC is unable to handle an excitation during a finite time. Having sin- gle electrons delivered at regular (ordered) time intervals (with narrower fluctuations of waiting times) as opposed to randomly (Poisson-like process) could avoid wasting excitations during such overly long blocking periods. The experimental implementation of our proposal as- sumes PSIIRC units that have been modified to have no quinones as has been done for the protein samples used in Ref.29. This will ensure that attachments of electrodes to individual PSIIRC units are at the level of the elec- tron donor and electron acceptor pigments. We envision metal-protein junctions and a scanning tip microscopy setup as those that have been realised for photosystem I (PSI) units9 and for reaction centre-enriched purple bac- terial membranes4. Genetic manipulation of both the oxidizing and reducing sites could allow covalent attach- ment of the protein to the electrodes across which the photo-current could be measured9. To measure the ele- mentary transfer events from the electron acceptor site to the drain electrode we envision a device, such as a sili- con field-effect transistor56,57, capable of detecting single charges and feasible to be integrated in the scanning tip setup at room temperature as shown in Fig. 1(a). The main limitation of isolated natural photosynthetic proteins for realistic, long-lived photovoltaic applications lies in the photodamage they experience. In PSIIRC this occurs in the D1 protein resulting in a lifetime as short as tens of minutes58,59. Emerging organic alternatives, such as the synthesis of man-made protein maquettes60 has opened the possibility of building nanometric units accurately mimicking the structure of photosynthetic sys- tems yet displaying enhanced photostability. Merging this area with photovoltaics may unleash an unforeseen remarkable development. 8 From the theoretical view point a few remarks must be made. As specified in the Methods section, we as- sume a simplified description of the coupling between the electronic system and the leads. This neglects the pos- sibility of the leads coupling to localized vibronic states rather than to bare electronic degrees of freedom. The same approximation has been used in other similar sys- tems with arbitrary electron-phonon couplings61,62 and has been shown to give relevant physical insight. The full extent of this effect in our system remains to be in- vestigated. There are also additional questions about the thermodynamic consistency of calculating the power out- put across with the phenomenologically modeled load as raised in Ref.63. It will therefore be important to inves- tigate entropy production64 for our model photocell to assess its consistency with the second law. Finally, our work has focused on the zero-frequency noise showing that, in this case, it is dominated by the population dynamics as confirmed by our comparison between the hybrid and the approximate frameworks. An extension of our study to investigate finite-frequency noise65 could therefore be a suitable alternative to ob- tain signatures of quantum coherence. More generally, current statistics measurements also potentially offer a non-invasive, single system level probe of charge trans- fer phenomena in a wide range of biological66,67 and chemical systems68,69 ranging from charge transfer along molecular wires made from DNA strands66 to general donor-bridge-acceptor systems68,69 or to unveil vibra- tional mechanisms for odour receptors67. METHODS Dynamical evolution of electronic excitations We consider an exciton dynamics described the basis of i corresponds to = Pi eiiihi + Pij Tij(iihj + jihi) by Hel where single- excitation states of the six core chromophores i.e. {PD1i, PD2i, ChlD1i, ChlD2i, PheD1i, PheD2i} and ei are onsite energies given in Table S1. The six eigenstates of Hel are denoted as X1i to X6i with corresponding eigenenergies EX1 to EX6 in ascending order. The electronic operators iihi couple linearly with coupling gi to identical baths of harmonic oscilla- k + bk). The strength of the system-bath interaction is quantified by the spectral density that will be of the form J(ω),JD(ω), J1(ω) or J2(ω). tors HI = Pi,k giiihi(b† To describe the full PSIIRC photocell dynamics under the operation conditions illustrated in Fig. 1 we consider two frameworks. (1) A hybrid framework that accounts for a non-perturbative approach to the exciton dynamics using the hierarchical equations of motion36 -- 38 in com- bination with incoherent transfer rates70 to and from all other states. The non-perturbative expansion of the exci- ton dynamics is used to account accurately for the effects of JD(ω), J1(ω) or J2(ω). Incoherent rates connecting the exciton states with the rest of states in the photocell are defined using a Lindblad dissipator coupled to each auxiliary density matrix in the expansion70. Supplemen- tary Note 2 presents further details of the hierarchical expansion of exciton dynamics under this scheme. Con- verged dynamics are obtained by terminating the hier- archical expansion at level N = 8 for JD(ω) and level N = 5 for J1(ω) and J2(ω). Only the K = 0 Matsubara term was explicitly accounted for, though a Markovian truncation term for Matsubara frequencies was included to capture some finite temperature effects37. (2) A Pauli master equation for electronic state pop- ulations is also considered, similarly to the approach fol- lowed on Ref.26. The Pauli rate equations have the form P ii = M P ii, where P ii is a vector of state popula- tions in the basis {gi, X1i, · · · X6i, Ii, αi, βi} and M is a stochastic matrix containing the rates for transfer between these electronic states. Modified Redfield theory as presented in Supplementary Note 2 is used to compute population transfer among exciton states45,71. In both frameworks (1) and (2) we assume weak and incoher- ent coupling from excited states to the primary CT state as well as weak coupling between charge transfer states. The transfer from exciton states Xni to the intermedi- ate CT Ii are given by Generalised Forster theory, and Forster-like rates are used to describe transfer between the CT states Ii and αi72. Other incoherent rates are described in Supplementary Note 2. Theoretical validity of this framework is discussed in Supplementary Note 3 along with a systematic comparison of the predictions of frameworks (1) and (2). Theory of full counting statistics We envisage our photocell positioned between source and drain leads which supply or remove electrons from the system respectively. The leads are taken as weakly coupled fermionic reservoirs in the limit of infinite bias, such that their influence is described by Lindblad-type dissipators33, as specified in Supplementary Note 5. With weak coupling to the leads, the theory of full count- ing statistics31,32,65 provides a framework to investigate the cumulants of the current passing through the sys- tem. This framework is applied to both the hybrid, non-perturbative approach, and the approximate Pauli model we use to describe the dynamics. For both mod- els a time-local master equation σ(t) = Mσ(t) can be 9 constructed, where the state of the system σ(t) is prop- agated through time by an operator M. This dynamical equation is augmented by a counting field χ, used to sin- gle out the incoherent transition across which the elec- tron statistics is counted. For our photocell this is the transition from state αi to state βi where an electron is transferred to the drain lead. This leads to the time propagator M(χ) = M0 + eiχMJ where M0 describes the time evolution of the system between counting events and MJ is the jump matrix describing hopping events between the system and the drain lead. Further details on the calculation of non-perturbative electron counting statistics using the hierarchical equations of motion47 are given in Supplementary Note 5. The zero-frequency cu- mulants are encoded in the probability distribution of the number of electrons that hop into the drain lead in some long time period32. A recursive scheme is then fol- lowed which generates zero-frequency current cumulants up to any order46 and expresses them in terms of the jump matrix MJ and the pseudo-inverse R of the time propagator. The mean hI 1i and noise hI 2i are given by hI 1i = hh0MJ 0ii hI 2i = hh0MJ − 2MJ RMJ 0ii (7) (8) where hh0 and 0ii are the left and right steady state eigenvectors of the time propagator. ACKNOWLEDGEMENTS The authors would like to thank Clive Emary, Elisabet Romero, Vladimir Novoredezkin and Jeroem Elzerman for helpful discussions. Financial support from the Engi- neering and Physical Sciences Research Council (EPSRC UK) Grant EP/G005222/1 and from the EU FP7 Project PAPETS (GA 323901) is gratefully acknowledged. AUTHOR CONTRIBUTIONS STATEMENT A.O-C designed the research, R.S and H.H-N carried out the calculations. RS, H.H-N, RvG and A.O-C anal- ysed the results and wrote the manuscript. ADDITIONAL INFORMATION The authors declare no competing financial interests. ∗ [email protected] 1 R. Blankenship, Molecular Mechanisms of Photosynthesis, Wiley-Blackwell, 2001. 2 R. Blankenship, D. Tiede, J. Barber, G. Brudvig, G. Flem- ing, M. Ghirardi, M. Gunner, W. Junge, D. Kramer, A. Melis, T. Moore, C. Moser, D. Nocera, A. Nozik, D. Ort, W. Parson, R. Prince and R. Sayre, Science, 2011, 332, 805 -- 9. 3 E. Wientjes, H. van Amerongen and R. Croce, Nat. Com- mun., 2013, 117, 11200 -- 11208. 4 M. Kamran, V. Friebe, J. Delgado, T. Aartsma, R. Frese and M. Jones, Nat. Commun., 2015, 6, 10 DOI:10.1038/ncomms7530. 35 F. Herrera and F. Spano, Phys. Rev. Lett., 2016, 116, 5 O. Yehezkeli, R. Tel-Vered, D. Michaeli, I. Willner and 238301. R. Nechushtai, Photosyn. Res., 2014, 120, 71 -- 85. 6 M. Gratzal, Nature, 2001, 414, 338 -- 344. 7 B. Reiss, D. Hanson and M. Firestone, Biotechnol. Prog., 2007, 23, 985 -- 989. 8 T. Mikayama, T. Miyashita, K. Iida, Y. Suemori and M. Nango, Mol. Cryst. Liquid Cryst., 2006, 445, 291 -- 296. 9 D. Gerster, J. Reichert, H. Bi, J. Barth, S. Kaniber, A. Holleitner, I. Visoly-Fisher, S. Sergani and I. Carmeli, Nat. Nanotech., 2012, 7, 673 -- 6. 10 Y. Blanter and M. Buttiker, Phys. Rep., 2000, 336, 1 -- 166. 11 G. Kiesslich, E. Scholl, T. Brandes, F. Hohls and R. Haug, Phys. Rev. Lett., 2007, 99, 206602. 12 N. Ubbelohde, C. Fricke, C. Flindt, F. Hohls and R. Haug, Nat. Commun., 2012, 3, 612. 13 F. Haupt, F. Cavaliere, R. Fazio and M. Sassetti, Phys. Rev. B, 2006, 74, 205328. 14 W. Belzig, Phys. Rev. B, 2005, 71, 1 -- 4. 15 H. Hossein-Nejad, A. Olaya-Castro, F. Fassioli and G. Sc- holes, New J. Phys., 2013, 15, 083056. 16 A. W. Rutherford and P. Faller, Philosophical Transac- tions of the Royal Society of London B: Biological Sciences, 2003, 358, 245 -- 253. 17 M. Groot, N. Pawlowicz, L. van Wilderen, J. Breton, I. van Stokkum and R. van Grondelle, Proc. Nat. Acad. Sci., 2005, 102, 13087 -- 13092. 18 A. Holzwarth, M. Muller, M. Reus, M. Nowaczyk, J. Sander and M. Rogner, Proc. Nat. Acad. Sci., 2006, 103, 6895 -- 6900. 19 E. Romero, I. van Stokkum, V. Novoderezhkin and J. Dekker, Biochem., 2010, 49, 4300 -- 4307. 20 T. Renger and E. Schlodder, J. Photochem. Photobio. B, 2011, 104, 126 -- 41. 21 V. Novoderezhkin, E. Andrizhiyevskaya, J. Dekker and 36 A. Ishizaki and G. Fleming, J. Chem. Phys., 2009, 130, 234111. 37 Q. Shi, L. Chen, G. Nan, R. Xu and Y. Yan, J. Chem. Phys., 2009, 130, 084105. 38 Y. Tanimura, J. Chem. Phys., 2012, 137, 22A550. 39 K. Dorfman, D. Voronine, S. Mukamel and M. Scully, Proc. Nat. Acad. Sci., 2013, 110, DOI:10.1073/pnas.1212666110. 40 C. Creatore, M. Parker, S. Emmott and A. Chin, Phys. Rev. Lett., 2013, 111, 253601. 41 W. Shockley and H. Queisser, J. Appl. Phys., 1961, 32, 510. 42 R. Ross and M. Calvin, Biophys. J., 1967, 7, 595 -- 614. 43 N. Killoran, S. Huelga and M. Plenio, J. Chem. Phys., 2015, 143, 155102. 44 T. Forster, Modern Quantum Chemistry, Part III. B. Ac- tion of Light and Organic Crystals., Academic Press, New York, 1965, pp. 93 -- 137. 45 M. Yang and G. Fleming, Chem. Phys., 2002, 275, 355 -- 372. 46 C. Flindt, T. Novotn´y, A. Braggio and A.-P. Jauho, Phys. Rev. B, 2010, 82, 155407. 47 R. Stones and A. Olaya-Castro, 2017, arXiv:1705.02320. 48 U. Fano, Phys. Rev., 1947, 72, 26 -- 29. 49 J. Egues, S. Hershfield and J. Wilkins, Phys. Rev. B, 1994, 49, 13517 -- 13527. 50 F. Fuller, J. Pan, A. Gelzinis, V. Butkus, S. Seckin Senlik, D. Wilcox, C. Yocum, D. Valkunas, L. Abramavicius and J. Ogilvie, Nat. Chem., 2014, 6, 706 -- 711. 51 E. Romero, V. Novoderezhkin and R. van Grondelle, Na- ture, 2017, 543, 355 -- 365. 52 L. Tsimring, Rep. Prog. Phys., 2014, 77, 026601. 53 J. Raser and E. O'Shea, Science, 2005, 309, 2010 -- 2013. 54 R. Brunetti, F. Affinito, C. Jacoboni, E. Piccinini and R. van Grondelle, Biophys. J., 2005, 89, 1464 -- 81. M. Rudan, J. Comp. Elec., 2007, 6, 391 -- 394. 22 V. Novoderezhkin, A. Yakovlev, R. van Grondelle and 55 L. Cardelli, A. Csik´asz-Nagy, N. Dalchau, M. Tribastone V. Shuvalov, J. Phys. Chem. B, 2004, 108, 7445 -- 7457. and M. Tschaikowski, Scientific Reports, 2016, 6, 20214. 23 Y. Umena, K. Kawakami, J. Shen and N. Kamiya, Nature, 56 K. Nishiguchi and A. Fujiwara, Nanotechnology, 2009, 20, 2011, 473, 55 -- 60. 175201. 24 B. Diner and F. Rappaport, Ann. Rev. Plant Bio., 2002, 57 K. Nishiguchi, Y. Ono and A. Fujiwara, Appl. Phys. Lett., 53, 551 -- 580. 25 M. Steffen, K. Lao and S. Boxer, Science, 1994, 264, 810 -- 816. 26 V. Novoderezhkin, E. Romero, J. Dekker and R. van Gron- delle, ChemPhysChem, 2011, 12, 681 -- 8. 2011, 98, 193502. 58 K. Brinkert, F. Le Formal, X. Li, J. Durrant, A. Rutherford and A. Fantuzzia, Biochim. Biophys. Acta., 2016, 1857, 1497 -- 1505. 59 M. Edelman and A. Mattoo, Photosynth. Res., 2008, 98, 27 G. Raszewski, B. Diner, E. Schlodder and T. Renger, Bio- 609 -- 20. physical Journal, 2008, 95, 105 -- 119. 28 V. Novoderezhkin, J. Dekker and R. van Grondelle, Bio- phyical Journal, 2007, 93, 1293 -- 1311. 29 E. Romero, R. Augulis, V. Novoderezhkin, M. Ferretti, J. Thieme, D. Zigmantas and R. van Grondelle, Nat. Phys., 2014, 10, DOI:10.1038/nphys3017. 30 P. Peterman, H. van Amerongen, R. van Grondelle and 60 B. Lichtenstein, C. Bialas, J. Cerda, B. Fry, P. Dutton and C. Moser, Angew Chem. Int. Ed. Engl., 2015, 54, 13626 -- 9. 61 A. Braggio, C. Flindt and T. Novotn´y, J. Stat. Mech., 2009, 2009, P01048. 62 D. Santamore, N. Lambert and F. Nori, PRB, 2013, 87, 075422. 63 D. Gelbwaser-Klimovsky and A. Aspuru-Guzik, Chem. J. Dekker, Proc. Natl. Acad. Sci., 1998, 95, 6128 -- 6133. Sci., 2017, 8, 1008. 31 D. Bagrets and Y. Nazarov, Phys. Rev. B, 2003, 67, 64 M. Esposito, K. Lindenberg and C. Van den Broeck, New 085316. J. Phys., 2010, 12, 013013. 32 D. Marcos, C. Emary, T. Brandes and R. Aguado, New J. 65 C. Emary, D. Marcos, R. Aguado and T. Brandes, Phys. Phys., 2010, 12, 123009. Rev. B, 2007, 76, 161404. 33 U. Harbola, M. Esposito and S. Mukamel, Phys. Rev. B, 66 L. Xiang, J. Palma, C. Bruot, V. Mujica, M. Ratner and 2006, 74, 235309. 34 D. Coles, Y. Yang, Y. Wang, R. Grant, R. Tay- lor, S. Saikin, A. Aspuru-Guzik, D. Lidzey, J. Kuo- Hsiang Tang and J. Smith, Nat. Commun., 2014, 5, 5561. N. Tao, Nat. Chem., 2015, 7, 221 -- 226. 67 M. Franco, L. Turin, A. Mershin and E. Skoulakis, Proc. Nat. Acad. Sci., 2011, 108, 3797 -- 3802. 68 M. Delor, I. Sazanovich, M. Torie and J. Weinstein, Acc. 70 C. Kreisbeck, T. Kramer, M. Rodr´ıguez and B. Hein, J. Chem. Res., 2015, 48, 1131 -- 1139. 69 A. Bakulin, R. Lovrincic, X. Yu, O. Selig, H. Bakker, Y. Rezus, P. Nayak, A. Fonari, V. Coropceanu, J. Bredas and D. Cahen, Nat. Commun., 2015, 6, DOI:10.1038/ncomms8880. Chem. Theory Comput., 2011, 7, 2166 -- 2174. 71 W. M. Zhang, T. Meier, V. Chernyak and S. Mukamel, J. Chem. Phys., 1998, 108, 7763 -- 7774. 72 E. O'Reilly, Ph.D. thesis, University College London, 2014. 11
1806.08373
1
1806
2018-06-21T18:07:14
Coagulation model for ion channels in the lipid membranes
[ "physics.bio-ph", "q-bio.SC" ]
A two-dimensional (2D) model is developed to describe the growth of spotty patterns (domains) in the lipid membranes which result from the coagulation of ion channels. It is assumed that the ion channels can coagulate when activated, e.g. the neurotransmitter may act as the surface-active substance for some ion channels in the lipid matrix or the activation may result in increase of polarity of the ion channels that give rise to their coagulation process due to the Brownian diffusive motion. The results show that the domain radius rD of the ion channels scales with time as rD~time^0.4.The slowing and subsequent saturation of the domain radius at longer time periods are in qualitative agreement with experiments.
physics.bio-ph
physics
Coagulation model for ion channels in the lipid membranes I.A. Larionov *, A.L. Larionov *, E.E. Nikolsky †‡, M.N. Shneider § *Institute of Physics, Kazan Federal University, Kremlevskaya str. 18, Kazan 420008, Russia †Kazan Institute of Biochemistry and Biophysics, Russian Academy of Sciences, Kazan 420008, Russia ‡Kazan State Medical University, Butlerov str. 49, Kazan, 420012 Russia §MAE Department, Princeton University, Princeton, NJ 08544, USA ABSTRACT A two-dimensional (2D) model is developed to describe the growth of spotty patterns (domains) in the lipid membranes which result from the coagulation of ion channels. It is assumed that the ion channels can coagulate when activated, e.g. the neurotransmitter may act as the surface-active substance for some ion channels in the lipid matrix or the activation may result in increase of polarity of the ion channels that give rise to their coagulation process due to the Brownian diffusive motion. The results show that the domain radius rD of the ion channels scales with time as rD~time 0.4 .The slowing and subsequent saturation of the domain radius at longer time periods are in qualitative agreement with experiments. Address reprint requests and inquiries: [email protected] It is known that transmembrane protein channels are not fixed in the membrane, but are subject to free lateral diffusion [1-5]. In the postsynaptic membrane, one quantum of the neurotransmitter activates about 1000 ion channels [6,7], and they can assemble into stable domains (clusters) with a number of channels in the domain up to several hundreds. The characteristic time for the domains formation and their decay is of the order of a few seconds. The domain formation is studied in experiments, see, e.g., [8], however, the qualitative and quantitative theory of this interesting and important phenomenon is still in the focus of discussion [8-14]. In the present paper, we show that the domain formation process can be a manifestation of Brownian coagulation and give quantitative characteristics of the growth process of domains for typical active lipid membranes. We assume that the activation of ion channels acquire the ability to form bonds and stick together as a result of collision. For ligand-activated channels, this can be caused, for example, by the action of the neurotransmitter, including as a surfactant. Activation of the ion channels may also result in increase of their polarity [14]. This leads to interaction (attraction) of the ion channels with force that is known to decrease rapidly with distance. spotty patterns Therefore, Randomly "collided" channels are stuck together. The resulting domain has a large lateral surface and, hence, lower mobility. of transmembrane channels formed accidentally during the initial stages of the membrane activity can be considered as nearly static initial nuclei and growth centers of the gigantic spotty clusters (domains) observed in experiments. A characteristic estimate of the formation time of the embryonic domain as a result of the Brownian motion of the transmembrane channels, which are subject to free lateral diffusion, gives , 1 (cid:2028)(cid:3036)~(cid:1864)(cid:2868)(cid:2870)/(cid:1830)(cid:3404)10(cid:2879)(cid:2871) s. Here l0~10nm is the average initial distance (spacing) between the outer surfaces of the ion channels, which have an outer radius of R1 =4 nm [15] and a typical value of the lateral diffusion coefficient is (cid:1830)(cid:3404)10(cid:2879)(cid:2869)(cid:2871) m2/s [1-5]. We the possibility of activated will assume that single activated ion channels come to our arbitrarily chosen "clustering center" from a region with a radius equal to or less than 0.5R2 [6,16,17]. This bound eliminates ion channels diffusing across the circle with radius R2 1 m in the time interval considered. The R2 boundary is set by the knowledge that one quantum of the neurotransmitter activates ion channels in a region on the order of 1 m2 (see, e.g., [6,16,17] and Fig. 1). The process of rapid domain size growth will end after a time interval of the order of 2 D R 2~  / 4 D  2.5 s. As no new channels are generated, the number and density of free channels in the membrane is reduced and the domain growth slows down and subsequently, ceases. to reduces The problem of the clustering of freely diffusing transmembrane channels the Brownian coagulation equation described by Smoluchowski's diffusion equation [18]. In this letter we will adopt for ion channels a coagulation model in 2D geometry. The 2D  n x y t , diffusion equation for the ion channels density is  n x y t , , y 2  Or, for isotropic case in cylindrical coordinates  n x y t , , x 2   n x y t ,  , (1) , t  ,        D   2  2       r  r where in Cartesian coordinates  n r t ,  t  D r   r   2   n r t , r  2  y x  2 , (2)    . Here the uniform sparse initial ion channel density (cid:1858)(cid:4666)(cid:2022)(cid:4667)(cid:3404)(cid:1866)(cid:2868)(cid:3404)(cid:1855)(cid:1867)(cid:1866)(cid:1871)(cid:1872), and the second boundary condition function g(cid:2870)(cid:4666)(cid:2028)(cid:4667) may depend on the model under study 0(cid:3409)g(cid:2870)(cid:4666)(cid:2028)(cid:4667)(cid:3409)(cid:1866)(cid:2868). In general g(cid:2870)(cid:4666)(cid:2028)(cid:4667) may be a decreasing function of time (cid:2028) in the form g(cid:2870)(cid:4666)(cid:2028)(cid:4667)(cid:3404)(cid:1866)(cid:2868)(cid:1857)(cid:1876)(cid:1868)(cid:4666)(cid:3398)(cid:2018)(cid:2028)(cid:4667) since the simplest case with g(cid:2870)(cid:4666)(cid:2028)(cid:4667)(cid:3404)0, i.e. the contribution from there are no newly created ion channels in the vicinity of the coagulation center. For simplicity we will consider only the first term of Eq. (3). This means that no activated ion channels cross the cirle with radius R2 within the time periods considered.  , n r t r  , (4)    j t D    The ion channels flux on the domain perimeter is [18] where (cid:1870)(cid:3005) is the domain (coagulation) radius. The total (cid:1870)(cid:3404)(cid:1870)(cid:3005)(cid:3404)(cid:1844)(cid:2869)√(cid:1840) per time, i.e. the rate of the coagulation number of ion channels crossing the domain perimeter (cid:1839)(cid:4666)(cid:1872)(cid:4667)(cid:3404)(cid:3515)(cid:1862)(cid:4666)(cid:1872)(cid:4667)(cid:1856)(cid:1838)(cid:3404)(cid:1862)(cid:4666)(cid:1872)(cid:4667)∙2(cid:2024)(cid:1844)(cid:2869)√(cid:1840) events is then given by    Dr r  and the number of ion channels in the domain is time  0 time  0 N  M t dt    2  R N 1   j t dt . (5)  time N We have to take into account the relative independent diffusive motion of the domain and single ion channels where the domain diffusion coefficient scales with its radius as (cid:1830)(cid:3005)∝(cid:1870)(cid:3005)(cid:2879)(cid:2869) (see, e.g., [2,8]) at least in the large (cid:1870)(cid:3005) limit. Therefore we replace D with Deff, where Deff is given by (cid:1830)(cid:3032)(cid:3033)(cid:3033)(cid:3404)(cid:1830)(cid:3397)(cid:1830)(cid:3005)(cid:3404)(cid:1830)(cid:3436)1(cid:3397) 1√(cid:1840)(cid:3440) This gives us the equation for the number of ion channels in the domain and the domain (coagulation) radius (cid:1870)(cid:3005)(cid:3404)(cid:1844)(cid:2869)√(cid:1840). The numerical value of R1 determines the maximum density of ion channels it corresponds to 18 000 ion channels/m2. the domain. For R1 =4 nm [15]  n r t , r  , (6)    R D 1 1 N 2  We solved Eq. (6) by the least squares method and the results of the calculations are shown in Fig. 2 with the following parameters set: n0=2 000/m2 [17], R1 =4 nm, R2 =1000 nm, and g2()=0. The values of the lateral diffusion coefficients are D=10-12 m2/s, and D=10-14 respectively. The calculations show that the number of ion channels in the central domain N and the domain radius (cid:1870)(cid:3005) scale with time m2/s, D=10-13 m2/s, right, from left to in          dt r r  D   1  0 and the diffusion coefficient D as 6  2.6 10 N   0.5 ; Dr R 1  D time  0.4 . The relative independent diffusive motion of the domain and single ion channels affects the domain radius growth 2 where , ,  t   2  R 2 2 1  G r with   k 1  2  k J  2 0 J  s   k   2   k 0   r  k k  s J 2  k 0      exp    D t 2  k R 2 1    , FIGURE 1. Sketch of the ion channels layout at the initial (a) and the final moment (b) of the domain growth process. Filled and open circles show the activated and inactivated ion channels, respectively. The inactivated ion channels (open circles) are beyond our consideration. The actual number of activated ion channels in the calculations is 0.25(cid:1866)(cid:2868)(cid:1844)(cid:2870)(cid:2870)(cid:3406) 1570. domain. With the initial condition (cid:1866)(cid:4666)(cid:1870),(cid:1872)(cid:3404)0(cid:4667)(cid:3404)(cid:1858)(cid:4666)(cid:1870)(cid:4667), and the boundary conditions for (cid:1872)(cid:3408)0, (cid:1866)(cid:4666)(cid:1870)(cid:3404)(cid:1844)(cid:2869),(cid:1872)(cid:4667)(cid:3404)0, and (cid:1866)(cid:4666)(cid:1870)(cid:3404)(cid:1844)(cid:2870),(cid:1872)(cid:4667)(cid:3404)g(cid:2870)(cid:4666)(cid:1872)(cid:4667), the solution has the form [19]: (cid:1866)(cid:4666)(cid:1870),(cid:1872)(cid:4667)(cid:3404)(cid:1516) (cid:1858)(cid:4666)(cid:2022)(cid:4667)(cid:1833)(cid:4666)(cid:1870),(cid:2022),(cid:1872)(cid:4667)(cid:1856)(cid:2022)(cid:3398)(cid:1830)(cid:1516)g(cid:2870)(cid:4666)(cid:2028)(cid:4667)Λ(cid:2870)(cid:4666)(cid:1870),(cid:1872)(cid:3398)(cid:2028)(cid:4667)(cid:1856)(cid:2028) For the 2D (or the 1D in cylindrical coordinates) case, in analogy with initial and boundary conditions for 3D case, we use the solution [19] for the density of particles in the (cid:3047)(cid:2868) (cid:3019)(cid:3118)(cid:3019)(cid:3117) , (3)  k   r  Y 0   k  J 0  k Y 0 ,    r  k R 1    0 r  J  k R 1    0J  and         2 /  , where s R R  1 functions, k - are the positive solutions of the equation: J 0Y  are the Bessel   s    , 0  Y s 0     Y 0  J    0 0  2  r t ,       G r , ,  t  .   R 2 and factor only a little except at the initial moments, that are beyond our interest. The number of ion channels in the domain is finally N520; this means that around 1/3 of the single ion channels, according to our estimate, form the central large domain. The ion channel layout in the domain formed by the coagulation processes due to the Brownian diffusive motion is in the disordered form. The increase of polarity of the activated ion channels may also affect their layout in the domain as reported in [14]. Investigation of the polarity or charge distribution within the activated ion channels are higly desirable both experimentally and theoretically. FIGURE 2. The domain radius Dr (or N0.5, where N is the total number of ion channels in the domain) versus time (solid lines). The values of the lateral diffusion coefficients m2/s, from left are D=10-12 to right, lines show N0.5=rD/R1=2.6×106×(D×time)0.4. m2/s, D=10-13 m2/s, and D=10-14 respectively. The dash-dotted In this Letter we showed that giant domain-spots, consisting of transmembrane protein channels observed in experiments with active membranes can form as a result of Brownian coagulation. In this case, the characteristic domain formation time is of the order of fractions of a second to a few seconds, depending on the value of the lateral diffusion coefficient. This is in agreement with known experimental data. It should be expected that at the end of the active phase of excitation (when, in our opinion, the binding mechanism ceases to function), the diffusion effect leads to smearing the domain and relaxation to a uniform distribution of the transmembrane channels occurs approximately within the same times. AUTHOR CONTRIBUTIONS M.N.S. proposed the model, I.A.L. and A.L.L. performed the calculations, E.E.N. initiated and supervised the work. All authors jointly wrote the article. 3 REFERENCES 1. Saffman, P.G., M. Delbrück, 1975. Brownian motion in biological membranes. Proc. Natl. Acad. Sci. USA. 72:3111–3113. 2. Gambin, Y., R. Lopez-Esparza, …, W. Urbach. 2006. Lateral mobility of proteins in liquid membranes revisited. Proc. Natl. Acad. Sci. USA. 103:2098–2102. 3. Bannai, H., S. Levi, …, A. Triller. 2006. Imaging the lateral diffusion of membrane molecules with quantum dots. 1:2628. doi:10.1038/nprot.2006.429 Protocols. 4. Almeida, P.F.F., W.L.C. Vaz. 1995. Lateral Diffusion in Membranes, in Handbook of biological physics, Vol. 1, Chapter 6, pp. 305-357, R. Lipowsky and E. Sackmann (eds.), Elsevier/North Holland, Amsterdam. 5. Ondrus, A.E., H.-l. D. Lee, …, J. Du Bois. 2012. Fluorescent saxitoxins for live cell imaging of single voltage-gated sodium ion channels beyond the optical diffraction 19:902–912. doi:10.1016/j.chembiol.2012.05.021 Chem. limit. Biol. Nat. 6. Kuffler, S.W., and D. Yoshikami. 1975. The number of transmitter molecules in a quantum: An estimate from iontophoretic application of acetylcholine at the neuromuscular synapse. J. Physiol. 251:465–482. 7. Anderson, C.R., and C.F. Stevens. 1973. Voltage clamp analysis of acetylcholine produced by end-plate current fluctuations at frog neuromuscular junction. J. Physiol. 235:655–691. 8. Stanich, C.A., A.R. Honerkamp-Smith, …, S.L. Keller. 2013. Coarsening Dynamics of Domains in Lipid Membranes. Biophys. J. 105:444-454. 9. Lifshitz, I. M., and V. V. Slyozov. 1961. The kinetics of precipitation from supersaturated solid solutions. J. Phys. Chem. Solids. 19:35–50. 10. Hughes, B.D., B.A. Pailthorpe, L.R. White. 1981. The translational and rotational drag on a cylinder moving in a membrane. J. Fluid Mech. 110:349–372. 11. Heuser, J.E., and S.R. Salpeter. 1979. Organization of acetylcholine receptors in quick-frozen, deep-etched, and rotary-replicated Torpedo postsynaptic membrane. J. Cell Biol. 82:150–173. 12. Unwin, N. 2013. Nicotinic acetylcholine receptor and the structural basis of neuromuscular transmission: from Torpedo postsynaptic membranes. insights Quarterly Reviews of Biophysics 46:283–322. 13. Vivas, O., C.M. Moreno, ..., B. Hille. 2017. Proximal clustering between BK and CaV1.3 channels promotes functional coupling and BK channel activation at low voltage. eLife 6:e28029. DOI: 10.7554/eLife.28029. 14. Unwin, N., and Y. Fujiyoshi. 2012. Gating Movement of Acetylcholine Receptor Caught by Plunge-Freezing. J. Mol. Biol. 422:617–634. 15. Adams, P.R. 1981. Acetylcholine receptor kinetics. J. Membr. Biol. 58:161–174. 16. Land, B.R., E.E. Salpeter, M.M. Salpeter. 1981. Kinetic parameters for acetylcholine intact neuromuscular junction. Proc. Natl. Acad. Sci. USA. 78:7200-7204. interaction in 17. Matthews-Bellinger, J., and M.M. Salpeter. 1978. Distribution of acetylcholine frog neuromuscular junctions with a discussion of some physiological implications. J. Physiol. 279:197–213. receptors at 18. Levich, V.G. 1962. Physicochemical hydrodynamics, Englewood Cliffs, N.J., Prentice-Hall. 700 p. 19. Polyanin, A.D. 2002. Handbook of Linear Partial Differential Equations for Engineers and Scientists, 1st Ed., Chapman and Hall/CRC; Taylor & Francis Group, Boca Raton, London, New York, Washington D.C. 800 p. 4
1711.02172
2
1711
2019-08-22T19:47:40
Active regeneration unites high- and low-temperature features in cooperative self-assembly
[ "physics.bio-ph", "cond-mat.stat-mech" ]
Cytoskeletal filaments are capable of self-assembly in the absence of externally supplied chemical energy, but the rapid turnover rates essential for their biological function require a constant flux of ATP or GTP hydrolysis. The same is true for two-dimensional protein assemblies employed in the formation of vesicles from cellular membranes, which rely on ATP-hydrolyzing enzymes to rapidly disassemble upon completion of the process. Recent observations suggest that the nucleolus, p granules and other three-dimensional membraneless organelles may also demand dissipation of chemical energy to maintain their fluidity. Cooperative binding plays a crucial role in the dynamics of these higher-dimensional structures, but is absent from classic models of 1-dimensional cytoskeletal assembly. In this Letter, we present a thermodynamically consistent model of actively regeneration with cooperative assembly, and compute the maximum turnover rate and minimum disassembly time as a function of the chemical driving force and the binding energy. We find that these driven structures resemble different equilibrium states above and below the nucleation barrier. In particular, we show that the maximal acceleration under large binding energies unites infinite-temperature local fluctuations with low-temperature nucleation kinetics.
physics.bio-ph
physics
Active regeneration unites high- and low-temperature features in cooperative self-assembly Physics of Living Systems Group, Massachusetts Institute of Technology, 400 Technology Square, Cambridge, MA 02139 Robert Marsland III∗ and Jeremy England (Dated: August 26, 2019) Cytoskeletal filaments are capable of self-assembly in the absence of externally supplied chemical energy, but the rapid turnover rates essential for their biological function require a constant flux of ATP or GTP hydrolysis. The same is true for two-dimensional protein assemblies employed in the formation of vesicles from cellular membranes, which rely on ATP-hydrolyzing enzymes to rapidly disassemble upon completion of the process. Recent observations suggest that the nucleolus, p granules and other three-dimensional membraneless organelles may also demand dissipation of chemical energy to maintain their fluidity. Cooperative binding plays a crucial role in the dynamics of these higher-dimensional structures, but is absent from classic models of 1-dimensional cytoskeletal assembly. In this Letter, we present a thermodynamically consistent model of actively regeneration with cooperative assembly, and compute the maximum turnover rate and minimum disassembly time as a function of the chemical driving force and the binding energy. We find that these driven structures resemble different equilibrium states above and below the nucleation barrier. In particular, we show that the maximal acceleration under large binding energies unites infinite-temperature local fluctuations with low-temperature nucleation kinetics. I. INTRODUCTION Kirschner and Mitchison pointed out in the 1980's that GTP hydrolysis in microtubules is responsible for the amazingly high rate of monomer exchange between fil- aments and the surrounding solvent [1]. An order-of- magnitude estimate using the measured association rates and binding energies shows that the dissociation rates in thermal equilibrium would be far too slow to support the massive structural rearrangements that take place over the course of the cell cycle. The large difference in chem- ical potential between the GTP and GDP pools in the cy- tosol allows the microtubule polymerization reaction to break detailed balance, speeding up the dynamics while maintaining the strength and stiffness demanded by the biological function of these structures. The coupling of nucleotide hydrolysis to monomer turnover seen in this particular case is in fact a generic feature shared by a variety of intracellular structures, which rely on similar mechanisms of active regeneration to enable timely responses to biochemical signals with- out sacrificing mechanical integrity. Recent studies have established the Hsp70 family of chaperone proteins as an all-purpose ATP-powered disassemblase, responsible for the rapid disassembly of disordered aggregates as well as of the protein coats that regulate vesicle formation in eukaryotic cells [2 -- 4]. Phase-separated intracellular droplets and granules appear to be fluidized by similar mechanisms [5, 6], and even the structure of interphase chromatin seems to be set by the competition between equilibrium phase separation and active disassembly [7]. ∗ Current address: Department of Physics, Boston University, 590 Commonwealth Avenue, Boston, MA 02215; Email: mars- [email protected] All these processes exhibit high levels of cooperativity: the dissociation rate of a given subunit from the structure depends on the presence or absence of multiple neigh- boring particles. This cooperativity gives rise to nonlin- earities that are not present in simple models of 1D cy- toskeletal filaments, which allow monomers to dissociate only at the ends of the filament where they have exactly one neighbor. These nonlinearities can make the equilib- rium state of cooperative systems relatively insensitive to changes in temperature, pH or other parameters except for a narrow range of values around a well-defined thresh- old [8, Ch. 9]. This robustness should allow low levels of active regeneration to accelerate the kinetics without significantly affecting the static properties of the struc- ture. But there is no general rule for determining how much acceleration can be tolerated, or what happens to the structure after this limit is reached. Several models have recently been developed that combine high-cooperativity equilibrium dynamics with a nonequilibrium driving force, leading to novel hypothe- ses about bistability in vesicle coat dynamics [9] and in the behavior of neurotransmitter protein receptors [10]. But these models are formulated in the limit of an infi- nite thermodynamic force, where at least one of the reac- tion steps is completely irreversible, and so they are not suitable for investigating the dependence of the system's properties on thermodynamic drive strength. In this work, we therefore present a fully reversible model of the intrinsically cooperative self-assembly of a high-dimensional structure, with active regeneration powered by a finite chemical potential difference ∆µ between chemical reservoirs. This model allows us to compute the kinetic acceleration at the disassem- bly threshold as a function of chemical potential dif- ference and monomer binding energy. We confirm that a small nonequilibrium driving force accelerates the ki- netics without dramatically modifying any static proper- ties. But at a critical value of ∆µ, the spectrum of fluc- tuations suddenly changes, with a corresponding jump in speed. We show that this novel phase combines the purely entropic local fluctuation spectrum of an infinite- temperature equilibrium state with the nucleation barrier that would be expected at the actual temperature. II. MODEL FORMULATION I = cI k and koff Our model describes a solution of proteins that can ex- ist in two different conformational states with identical internal free energy, illustrated by circles and squares in Figure 1: an "active" form that can bind to other active proteins to generate a larger structure, and an "inactive" form with no binding ability. Inactive proteins in solu- tion at concentration cI stochastically enter and leave free binding sites of an existing structure, with Poisson rates kon I = k per site, respectively. Since these proteins do not interact with each other or with the active proteins, neither rate is affected by the occu- pancy of other sites in the structure. For notational sim- plicity, we have chosen the units of concentration such that cI = 1 is the value at which these non-interacting proteins would occupy half of the available sites. Ac- tive proteins, at concentration cA, stick to other active proteins when they enter the structure, and the ratio of their association and dissociation rates is determined by the binding energy ∆G. These proteins bind to avail- able sites at a rate kon A = cAk per site, and dissociate at a rate koff A = k exp[−β∆G] where β = 1/(kBT ) is the inverse thermal energy scale. The binding energy is proportional to the fraction m ∈ [0, 1] of the N binding sites of the fully assembled structure that are currently occupied, so that ∆G = Jm for some constant J. This form of ∆G ignores the effect of spatial heterogeneity on the binding kinetics, but is commonly used in statistical physics to construct a qualitatively correct and analyti- cally tractable theory of phase transitions [11, Ch. 5]. The existence of two internal states with different bind- ing energies allows for a thermodynamic driving force to accelerate the kinetics. This acceleration is achieved by biasing the active-inactive transition in different direc- tions, depending on whether the protein is in solution or in the structure [12]. In solution, the transition should be biased towards the active state, to maintain a large pool of proteins ready to assemble. But in the structure, it should be biased towards the inactive state, so that bound proteins can be rapidly ejected and replaced. This cycle breaks detailed balance, and so is impossi- ble at thermal equilibrium. Cells power these cycles with nucleotide hydrolysis. In actin filaments, for example, the active monomer conformation is stabilized by bind- ing to ATP, and the inactive conformation is stabilized by binding to ADP [13]. When actin monomers are in solution, exchange of ADP for ATP takes place at a much faster rate than ATP hydrolysis, and the large concen- tration difference between the nucleotides under physio- 2 FIG. 1. Themodynamically consistent model of ac- tively regenerating cooperative self-assembly process. Color online. (a) Blue circles and red squares represent two distinct internal states of the assembling monomers with dif- ferent binding energies (∆G > 0 and 0, respectively). Arrows represent binding/unbinding and state-changing reactions. A steady-state probability current is maintained around this cy- cle by coupling the state change to ATP hydrolysis. The circles are stabilized by binding to ATP (small circles), and the squares by binding to ADP (diamonds). The free en- ergy of hydrolysis biases the state-change reaction of a bound circle towards the non-interacting square state. The square rapidly dissociates into the solution, where nucleotide ex- change (double-headed arrows) is much faster than hydrol- ysis, and transforms the particle back to the circle state. The particle can now bind to the structure again, completing the cycle. (b) Shaded curves are steady-state probability distri- butions pss(m) over the occupancy fraction m, for increas- ing values of the system size N , with ∆µ = 0, βJ = 8 and cA = 0.02. Black line is the effective free energy density, defined as f (m) = − limN→∞ pss(m)/N . logical conditions biases the transition towards the active ATP-bound state. But when a monomer is in a filament, hydrolysis is much faster than nucleotide exchange, and the large free energy of this reaction reverses the bias. Figure 1 illustrates how we implemented this general strategy within our model of cooperative assembly. We assume nucleotide is fast enough compared to hydrolysis for proteins in solution that the latter reaction is neg- ligible. The cytosolic concentrations cA and cI of the two conformations are thus fixed at the equilibrium val- ues determined by the ATP/ADP ratio. But within the structure, nucleotide exchange is forbidden. The transi- tion from the active to the inactive conformation takes place at a rate kI , and is necessarily coupled to ATP hy- drolysis. The reverse process then involves reversing the hydrolysis reaction, with the result that the rate kA for returning to the active state is much slower than it oth- erwise would be. Since the two conformations have the same intrinsic free energy, the free energy change during the transition comes only from the altered interactions of the protein with the rest of the structure, and from the nucleotide hydrolysis. Local detailed balance [14, 15] then requires that kI kA = eβ(∆µ0−∆G) (1) where ∆µ0 is the free energy released in the hydrolysis reaction, related to the full chemical potential difference kIkAATPADPm10(a)(b)f(m)pss(m)N=10N=50N=200cAkcIkkkeG ∆µ by ∆µ0 = ∆µ − kBT ln [ATP] [ADP] = ∆µ − kBT ln cA cI . (2) and the second equality results from rapid nucleotide ex- change in solution. To study kinetic acceleration in this model, we had to identify a fixed timescale to serve as a point of ref- erence. In many biochemical contexts, association rates are mainly determined by the speed of diffusion to the binding site, and are insensitive to changes in binding energy. We therefore chose 1/k as the basic timescale, and set k = 1 for the remainder of the analysis. A trivial way to accelerate the kinetics without break- ing detailed balance would be to increase cI or kI , so that most of the particles in the structure are inac- tive, and the typical dissociation rate increases from koff A = exp[−β∆G] to koff I = 1. But since the inactive particles are non-interacting, this would really represent a transient density fluctuation in a concentrated solution of freely diffusing particles, and not a process of self- assembly. We therefore restricted our attention to the regime cI , kI (cid:28) 1, so that inactive monomers dissociate from the lattice much faster than they are added from the solvent or created from active monomers in the structure. III. STEADY-STATE SOLUTION We proceeded to quantify the maximum steady-state turnover rate and the minimum time required for total disassembly under these constraints. To this end we ob- tained a closed set of dynamical equations for the struc- ture occupancy m by eliminating the short-lived states with bound inactive particles from the original model, as described in the Appendix [16]. In the resulting coarse- grained model, the occupancy can stochastically increase by an increment ∆m = 1/N in a Poisson process with rate w+(m) = N (1 − m)cA[1 + e−β∆µq(m)] and can decrease by 1/N with rate w−(m) = N me−βJm[1 + q(m)] where q(m) ≡ kI e−βJm + kI e−β∆µ0 (3) (4) (5) contains the contribution of transient visits to the inac- tive state. This contribution grows with increasing ∆µ0, as the transition to the inactive state becomes more and more irreversible. In the limit of large N , the steady-state distribution generated by these dynamics scales as pss(m) ∝ e−N f (m), 3 FIG. 2. Active regeneration accelerates turnover and disassembly. Color online. (a) Local turnover rate koff in assembled phase for four values of βJ. cA and ∆µ0 are tuned to the coexistence point, with cI = 0.0001. (b) Mean time τdiss for switching from high to low m states under the same conditions, for a structure of size N = 100. (c) Effective free energy landscapes f (m) on the βJ = 12 curve at the four values of ∆µ indicated by the dots with corresponding colors in panel (a). (d) Stochastic trajectories for N = 100 at the three nonzero ∆µ values, plotted in the same colors. and the function f (m) can be calculated analytically, as shown in the Appendix: (cid:2) −kI (e−β∆µ0 + e−β∆µ)eβJm(cid:3) (cid:18) −kI (e−β∆µ0 + 1)eβJm(cid:3)(cid:19) (cid:2) + C. (6) 1 βJ f (m) =feq(m) + Li2 − Li2 The first term is related to the equilibrium free energy for a fully coupled Ising model: feq(m) = m ln m + (1 − m) ln(1 − m) − m ln cA − βJ 2 m2 (7) and the second is expressed in terms of the dilogarithm function ∞(cid:88) k=1 Li2(z) ≡ zk k2 , (8) with a normalization constant C added at the end. The probability pss(m) converges in the N → ∞ limit to a delta function centered on the occupancy m∗ that minimizes f (m), as illustrated in Figure 1. For βJ > 4, f (m) generically exhibits (at least) two local minima, one at high m corresponding to an assembled phase, and one at low m corresponding to a disassembled phase. mt100⌧diss051015µJ=14ko↵051015µm100.0f(m)(b)(c)(d)(a)246⇥10710910211033104581012141061041020.30.6µ=06.59.510.5 IV. TURNOVER AND DISASSEMBLY SPEED In this limit, we can approximate the steady-state turnover rate per particle by evaluating w− at the global minimum m∗: koff ≡ w−(m∗) N m∗ ∗ = e−βJm [1 + q(m∗)]. (9) We computed the maximum value of koff in the assembled phase as a function of J and ∆µ, with kI = 0.01 and cI = 0.0001 held fixed to ensure they stay sufficiently small. This maximum occurs when ∆µ0 is tuned to the coexistence point, so that any further acceleration would send the system over the threshold into the disassembled phase. At ∆µ = 0, koff ≈ 2e−βJ is several orders of magnitude smaller than the rate for free diffusion out of a binding site. At the other extreme, ∆µ → ∞, every inactivation reaction ends in ejection from the lattice, so koff ≈ kI . Panel (a) of Figure 2 shows how the turnover rate crosses over from the former limit to the latter, for four different values of βJ spanning a physiologically relevant range of 5 to 8 kcal/mol at T = 300 K. koff increases rapidly for moderate values of β∆µ, while the qualitative shape of the effective free energy landscape remains the same, with narrow minima near m = 0 and m = 1, and a single barrier in between. Near ∆µ = 10kBT , however, the landscape changes drastically, and an additional lo- cal minimum emerges. When this local minimum be- comes the global minimum, koff discontinuously jumps to its high-∆µ limiting value. This discontinuity corre- sponds to a first-order phase transition in the N → ∞ limit. Notably, the physiological value of ∆µ ∼ 20kBT for ATP/ADP is more than sufficient to cross this tran- sition where it exists and reach the plateau for all four coupling strengths. The effect of active regeneration is magnified when we consider the mean first-passage time τdiss for the global transition from the assembled to the disassembled state, when the system is taken across the phase boundary by an infinitesimal change in one of the parameters [17]: ∗(cid:88) m 1(cid:88) τdiss = pss(m(cid:48)) m(cid:48)=m † pss(m)w−(m) ∗ )−f (m )]. m=m0 ∼ eN [f (m (10) (11) Here m† is the location of the largest value of f (m) be- tween its two local minima, and the second line represents the qualitative behavior of τdiss in the limit of large N . See the Appendix for derivations of these expressions. The factor of N in the exponent of Equation (11) can rapidly inflate τdiss to astronomical values. Panel (b) of Figure 2 shows how τdiss decreases as a function of ∆µ when N = 100, for the same four values of βJ as in panel (a). The disassembly time is mainly determined by the height of the nucleation barrier f (m†)− f (m∗), which re- mains unchanged at the nonequilibrium phase transition 4 observed above, and so the jump in τdiss at the transition point is barely visible on the plot. But the slope of the exponential decrease is extremely large, so that a small change in ∆µ can shift the disassembly transition from being effectively impossible to being observable on acces- sible timescales. This is illustrated in panel (d), where the shift in ∆µ from 9.5 to 10.5 kBT allows several spon- taneous assembly-disassembly transitions to take place over the plotted timespan. V. EFFECTIVE TEMPERATURES Having established that active regeneration can gen- erate substantial kinetic acceleration without disassem- bling the structure, we proceeded to investigate the ex- tent to which the fluctuations and dynamics of these ac- celerated states can be captured by an equilibrium model with modified parameters. Figure 3 compares the effec- tive free energy of the actively regenerating system on both sides of the nonequilibrium phase transition to sim- ilar equilibrium free energy landscapes. The low-m be- havior always agrees with feq(m) at the actual temper- ature, coupling and monomer concentration. But as m increases, f (m) smoothly transitions to a different equi- librium landscape with altered parameters. When ∆µ is below the phase transition, as in panel (a), this al- tered landscape is obtained by decreasing cA while keep- ing the temperature fixed, until the high-m local mini- mum agrees with the driven system, as shown in the Ap- pendix. This causes a small increase in the turnover rate for the equilibrium model to koff = 6.7×10−6, almost ten times less than the value of 6.3× 10−5 observed in Figure 2. The barrier to disassembly, however, is significantly lowered by this concentration shift, causing an exponen- tial drop in the disassembly time τdiss. The equilibrium model slightly overestimates the barrier reduction, pre- dicting a τdiss eight-fold smaller than the actual value in the plotted example. On the other side of the nonequilibrium phase tran- sition, the high-m part of the landscape is completely different. The matching equilibrium landscape has an effective temperature higher than the critical tempera- ture Tc = J/4kB, and only contains a single local mini- mum. When e−β∆µ0 (cid:28) e−βJm /kI , the dynamics near the steady state become independent of J, and the ef- fective temperature is infinite, as illustrated in panel (b) of Figure 3 and derived in the Appendix. But the ac- tual system still contains the nucleation barrier from the low-m regime, which now determines the disassembly ki- netics. ∗ The emergence of this high effective-temperature phase with a low-temperature nucleation barrier is a very ro- bust phenomenon, which should also arise in more com- plex models. The essential assumption is that the oper- ation of the regeneration process on a given particle is agnostic to the presence or absence of other particles in the structure. At high enough coupling, active regenera- 5 tem at the same temperature but lower monomer con- centration. In this case, we expect that the effective monomer concentration even in more detailed models may be predictable from fundamental thermodynamic bounds, following a recently developed method for an- alyzing nonequilibrium assembly processes under moder- ate drive strength [19]. In the limit of strong driving, we identified a novel nonequilibrium phase with infinite ef- fective temperature, which preserves the nucleation bar- rier from the original temperature. This is the regime we should find when regeneration is driven by nucleotide hydrolysis under physiological con- ditions, with ∆µ ∼ 20kBT . So far, a number of intracel- lular structures from liquid droplets [20] to much smaller protein aggregates [21] seem to be well-described by mod- els of equilibrium phase separation, but this is consistent with the predicted persistence of the equilibrium nucle- ation barrier in the strong driving regime. Testing our predictions will require further investigation of the statis- tics of the assembled phase. We expect that our analy- sis will be especially applicable to droplets whose con- densation is regulated by the phosphorylation state of the monomers [22], where a difference in phosphoryla- tion rates within and outside the droplet could set up a dissipative cycle analogous to the one we have described. Finally, although we designed our model to capture the physics of high-dimensional structures, our results may also have implications for the behavior of cytoskele- tal filaments. Several mechanisms have been proposed that add effective cooperativity to these 1-dimensional structures through length-dependent arrival of special- ized molecules that modify the dynamics [23]. A thermo- dynamic analysis of these systems raises additional chal- lenges beyond those produced by the intrinsic coopera- tivity of higher-dimensional models, because the energet- ics of the underlying molecular motor transport and en- zyme activity would also have to be explicitly accounted for. But an exploration of the connections between these models could provide interesting avenues for future work. ACKNOWLEDGMENTS RM acknowledges Government support under and awarded by DoD, Air Force Office of Scientific Re- search, National Defense Science and Engineering Grad- uate (NDSEG) Fellowship, 32 CFR 168a, and through NIH NIGMS grant 1R35GM119461. We thank Zachary Slepian, Jordan Horowitz, Kirill Korolev and Pankaj Mehta for helpful comments. FIG. 3. Driven steady state combines features of two equilibrium landscapes. Color online. (a) Effective free energy landscape with ∆µ = 6.5kBT , below the nonequilib- rium phase transition (black), compared with the equilibrium landscape at the same temperature/coupling βJ = 12 and monomer concentration cA = 8.1 × 10−3 (blue) as well as a modified equilibrium landscape with the same temperature but a smaller monomer concentration cA = 8.8 × 10−4 (red). (b) Same plots for ∆µ = 15.2kBT , except that the modified equilibrium landscape (red) is obtained by making the tem- perature infinite, and setting the (constant) off-rate equal to kI = 0.01. tion is the dominant pathway for particle removal in the assembled phase, and at high enough ∆µ0, the thermo- dynamically required dependence of the reverse rate on the coupling becomes irrelevant. In this regime, all bind- ing sites behave independently, and the statistics can be determined by purely entropic considerations. Below the nucleation barrier, on the other hand, active regenera- tion becomes irrelevant, as long as it is sufficiently slow compared to the spontaneous dissociation rate (which we guaranteed here by requiring kI (cid:28) 1). As we pointed out above, this assumption is necessary if assembly is to proceed at all, and does not limit the generality of the argument. VI. DISCUSSION We have constructed a minimal model of active re- generation in cooperative self-assembly, and have shown that it can significantly accelerate monomer turnover and assembly/disassembly transitions, while maintaining a sharp distinction between assembled and disassembled states. This confirms that chemically powered regenera- tion can produce the combination of structural resilience and rapid responsiveness characteristic of living systems [18]. When the chemical driving force was weak, we ob- served that the local spectrum of fluctuations was only slightly modified, corresponding to an equilibrium sys- [1] Marc Kirschner and Tim Mitchison. Beyond self- assembly: From microtubules to morphogenesis. Cell, 45:329, 1986. [2] Rui Sousa. Structural mechanisms of chaperone medi- ated protein disaggregation. Frontiers in Molecular Bio- sciences, 1:12, 2014. [3] Ramiro H Massol, Werner Boll, April M Griffin, and Tomas Kirchhausen. A burst of auxilin recruitment de- termines the onset of clathrin-coated vesicle uncoating. Proc. Natl. Acad. Sci., 103(27):10265 -- 10270, 2006. m100.0f(m)0.2(a)(b)m100.0f(m)0.20.20.4Actively regeneratingEquilibriumModified Equilibrium [4] Till Bocking, Fran¸cois Aguet, Stephen C Harrison, and Tomas Kirchhausen. Single-molecule analysis of a molec- ular disassemblase reveals the mechanism of Hsc70-driven clathrin uncoating. Nature Structural & Molecular Biol- ogy, 18(3):295 -- 301, 2011. [5] Triana Amen and Daniel Kaganovich. Dynamic droplets: the role of cytoplasmic inclusions in stress, function, and disease. Cellular and Molecular Life Sciences, 72:401, 2015. [6] Anthony A. Hyman, Christoph A. Weber, and Frank Julicher. Liquid-Liquid Phase Separation in Biology. An- nual Review of Cell and Developmental Biology, 30:39, 2014. [7] Johannes Nuebler, Geoffrey Fudenberg, Maxim Imakaev, Nezar Abdennur, and Leonid Mirny. Chromatin Organi- zation by an Interplay of Loop Extrusion and Compart- mental Segregation. bioRxiv, 196261. [8] Philip Nelson. Biological Phyiscs: Energy, Information, Life. W. H. Freeman and Co., New York, 2008. [9] L. Foret and P. Sens. Kinetic regulation of coated vesi- cle secretion. Proceedings of the National Academy of Sciences, 105:14763, 2008. [10] V. M. Burlakov, N. Emptage, A. Goriely, and P. C. Bressloff. Synaptic Bistability Due to Nucleation and Evaporation of Receptor Clusters. Physical Review Let- ters, 108:028101, 2012. [11] Mehran Kardar. Statistical Physics of Particles. Cam- bridge University Press, Cambridge, 2007. [12] Job Boekhoven, Wouter E Hendriksen, Ger J M Koper, Rienk Eelkema, and J. H. van Esch. Transient assembly of active materials fueled by a chemical reaction. Science, 349:1075, 2015. [13] E. Korn, Marie-France Carlier, and Dominique Pan- taloni. Actin polymerization and ATP hydrolysis. Sci- ence, 238:638, 1987. [14] Udo Seifert. Stochastic thermodynamics, fluctuation theorems and molecular machines. Rep. Prog. Phys., 75:126001, 2012. [15] Robert Marsland and Jeremy England. Limits of predic- tions in thermodynamic systems: a review. Reports on Progress in Physics, 81:016601, 2018. [16] Simone Pigolotti and Angelo Vulpiani. Coarse graining of master equations with fast and slow states. The Journal of Chemical Physics, 128:154114, 2008. [17] N. G. van Kampen. Stochastic Processes in Physics and Chemistry. Elsevier, Amsterdam, 3rd edition, 2007. [18] Sophie Dumont and Manu Prakash. Emergent mechanics of biological structures. Molecular Biology of the Cell, 25:3461, 2014. [19] Michael Nguyen and Suriyanarayanan Vaikuntanathan. Design principles for nonequilibrium self-assembly. Pro- ceedings of the National Academy of Sciences, 113:14231, 2016. [20] Clifford P. Brangwynne, Peter Tompa, and Rohit V. Pappu. Polymer physics of intracellular phase transi- tions. Nature Physics, 11(11):899 -- 904, 2015. [21] Arjun Narayanan, Anatoli B Meriin, Michael Y Sher- man, and Ibrahim I Cisse. A First Order Phase Transi- tion Underlies the Formation of Sub-Diffractive Protein Aggregates in Mammalian Cells. bioRxiv, 148395. [22] Pilong Li, Sudeep Banjade, Hui-Chun Cheng, Soyeon Kim, Baoyu Chen, Liang Guo, Marc Llaguno, Javoris V. Hollingsworth, David S. King, Salman F. Banani, Paul S. Russo, Qiu-Xing Jiang, B. Tracy Nixon, and Michael K. 6 Rosen. Phase transitions in the assembly of multivalent signalling proteins. Nature, 483:336 -- 340, 2012. [23] Lishibanya Mohapatra, Bruce L. Goode, Predrag Je- lenkovic, Rob Phillips, and Jane Kondev. Design Princi- ples of Length Control of Cytoskeletal Structures. Annual Review of Biophysics, 45:85, 2016. [24] Steven A. Bender, Carl M. and Orszag. Advanced Math- ematical Methods for Scientists and Engineers. Springer- Verlag New York, New York, 1999. APPENDIX This Appendix contains derivations of the following expressions: 1. coarse-grained rates w+ and w−, Equations (3)-(5) 2. effective free energy landscape f (m), Equations (6)-(7) 3. disassembly time τdiss, Equations (10)-(11). It also includes additional plots comparing effective free energy landscapes f (m) with equilibrium landscapes, and an analytic derivation of the infinite effective tem- perature of the novel high-∆µ phase. A. Coarse-Grained Rates Pigolotti et al. have described a systematic procedure for eliminating short-lived states from a Markov process, which exactly preserves the stationary state of the origi- nal process, and approximates the dynamics to arbitrary precision in the limit of vanishing lifetime of the fast states [16]. In this section, we give a heuristic motivation for this procedure, and describe how we applied it to our system. In the main text, we required that the inactive monomers bind so poorly that they make up a negligible percentage of the total structure occupancy at any given time, by imposing cI , kI (cid:28) 1, (12) in units of time where the dissociation rate of inactive particles is equal to 1. This guarantees that the incoming rates are much smaller than at least one of the outgoing rates, so that the steady state will have much more prob- ability in the states with no particle or a bound active particle in a given site than in states with an inactive particle there. Now the state of the system is entirely specified by the occupancy m, and there are two ways for a site to change state: either through direct association/dissociation of an active monomer from the solution, or by transiently pass- ing through the inactive form. Since the exit rate from the inactive state is much faster than the time scale of the dynamics of interest (set by the small inactivation rate kI ) we can approximate the second pathway as its own Markov process. The rate for dissociation of a given par- ticle via the inactive conformation is simply the product of the inactivation rate kI and the probability that the site ends up with a different occupancy after the next jump (instead of returning to its starting point). This probability is equal to the ratio of the dissociation rate of the inactive particle (which equals 1 in our units) to the total rate of disappearance of the inactive particle kA + 1, which includes the rate of return to the active state without leaving the structure. Likewise, the associ- ation rate through this pathway is the product of the rate cI for adding an inactive particle to the structure and the probability that it relaxes to the active state instead of returning to the solvent. Thus the rates for adding and removing active particles via the inactive state are given by 1 koff AI = kI kon AI = cI 1 + kA kA 1 + kA , (13) (14) in accordance with Equation (9) of [16]. It is important to note that this procedure preserves the thermodynamics of the original process. The entropy released into the environment in the original model when an active particle is removed via the inactive state is sim- ply the logarithm of the product of the rate ratios (cf. [14]): ∆S = kB ln kI kA 1 cI (15) The entropy released in the coarse-grained model is the log-ratio of the new rates, which gives the same value, since the denominators of the fractions cancel out: ∆S = kB ln koff AI kon AI = kB ln kI × 1 kA × cI . (16) The total rate w+(m) for the transition from occu- pancy m to m + 1/N is the sum of the rates of all pos- sible ways of accomplishing this transition: particles can be added to any of the N (1− m) free sites, and they can be added to each site by either of the two pathways. Sim- ilarly, the m to m − 1/N transition rate w−(m) includes the two removal rates for all N m particles currently in the structure: (cid:18) (cid:18) w+(m) = N (1 − m) cA + cI w−(m) = N m e−βJm + kI (cid:19) (cid:19) kA kA + 1 1 kA + 1 (17) , (18) We can now to express both rates in terms of the ther- modynamic quantities ∆µ and ∆µ0, using Equation (2) from the main text to obtain cA cI cAkI cI kA eβ∆µ = eβJm. eβ∆µ0 (19) (20) = Substituting for cI and kA in favor of ∆µ and ∆µ0, we 7 have: where w+(m) = N (1 − m)cA[1 + e−β∆µq(m)] w−(m) = N me−βJm[1 + q(m)] (21) (22) q(m) ≡ kI e−βJm + kI e−β∆µ0 . (23) These are the expressions given in Equations (3)-(5) of the main text. B. Effective Free Energy Landscape The coarse-grained dynamics are one-dimensional, with hard boundaries at m = 0 and m = 1, so they cannot support any steady currents. The average rate of jumps from m to m+1/N values has to equal the average rate of jumps from m + 1/N to m in order to keep the probability distribution p(m) stationary: w+(m)pss(m) = w−(m + 1/N )pss(m + 1/N ). (24) This detailed balance relation implies that the model could also describe an undriven system whose free en- ergy landscape is given by the logarithm of pss(m). But the functional dependence of these energies on m does not resemble any readily identifiable physical situation. To analyze the phase behavior of the model, we are interested in the thermodynamic limit where N → ∞. For this reason, we discussed the steady state distribution in terms of the effective free energy density f (m), defined by f (m) ≡ − lim N→∞ 1 N ln pss(m). (25) The derivative of this quantity is related to the ratio of rates via Equation (24): df dm ≡ lim N→∞ f (m + 1/N ) − f (m) 1/N pss(m) pss(m + 1/N ) w−(m + 1/N ) w+(m) = lim N→∞ ln = lim N→∞ ln = ln w−(m) w+(m) (26) (27) (28) (29) If we substitute in the expressions for w− and w+ ob- tained in the previous section, we find df dm = ln m − ln(1 − m) − βJm − ln cA + ln 1 + q . 1 + e−β∆µq (30) where the constant of integration C can be used to nor- malize the distribution. This is Equation (6) of the main text. ∆τm = We can now integrate both sides of this equation to find f (m) up to a constant of integration, which will be used to normalize the probability distribution pss(m). The integrals of the first four terms are straightforward, so that the equilibrium free energy density at ∆µ = 0 takes on the familiar form: feq(m) = m ln m + (1 − m) ln(1 − m) − m ln cA − βJ 2 m2 (31) as stated in Equation (7) of the main text. The integral of the final term has no elementary ex- pression, but can be evaluated in terms of the dilogarithm function Li2(x) ≡ ln(1 − x) dx x (cid:90) k=1 zk k2 = − ∞(cid:88) (cid:18) −kI (e−β∆µ0 + 1)eβJm(cid:3)(cid:19) (cid:2) 1 βJ Li2 (cid:2) −kI (e−β∆µ0 + e−β∆µ)eβJm(cid:3) + C. (33) to give f (m) =feq(m) + − Li2 C. Disassembly Time The cooperativity of this model generates a distinct threshold for disassembly, which becomes a sharp first- order phase transition as N → ∞. Close to the thresh- old, a small change in any of the parameters can shift the probability of finding the system in the assembled phase from nearly 1 to nearly 0. An important kinetic property of the system is the rate at which it switches to the disassembled phase after this parameter change. We quantified this rate using the mean first-passage time τdiss for the system to reach the low-m local minimum m0 of f (m), starting from the high-m local minimum m∗. Since the dynamics are continuous across the tran- sition, we can compute this quantity at the coexistence point where both these local minima have the same value of f (m), and use this as a good approximation for the disassembly time. We can derive an expression for τdiss in terms of the steady-state distribution pss(m), following [17, XII.2]. The argument starts from a self-consistency equation for the mean first-passage time τm,m0 from an arbitrary m to the disassembled state m0. We initialize the system at m, then wait for a short amount of time δt. Now the system could be at m+∆m with probability w+(m)δt, at m−∆m with probability w−(m)δt, or remain at m with proba- bility 1 − [w+(m) + w−(m)]δt (where ∆m = 1/N ). The probability of going more than one step away from m has (32) w+(m)∆τm+∆m − w−(m)∆τm + O(δt) = −1. (35) 8 a probability proportional to δt2. Since we have waited at time δt, the mean time to reach m0 is now equal to τm,m0 − δt. The new waiting time can also be calculated by averaging the mean first-passage times of the possi- ble current system states, weighted by their probabilities. These two ways of calculating the average remaining time need to give the same answer, yielding: τm,m0 − δt =τm+∆m,m0w+(m)δt + τm−∆m,m0 w−(m)δt + τm,m0(1 − [w+(m) + w−(m)]δt) + O(δt2). (34) This expression can be simplified by subtracting τm,m0 from both sides, and dividing by δt: where we have defined ∆τm ≡ τm,m0 − τm−∆m,m0. Now we can take the limit δt → 0 and drop the O(δt) term, obtaining a recursion relation for ∆τm. Since w+(1) = 0 in Equation (21) above, we have ∆τ1 = 1/w−(1), and can use Equation (35) to construct all the ∆τm's from this starting point. Each iteration multiplies the previous increment ∆τm+∆m by w+(m)/w−(m), and adds a new term 1/w−(m). Thus we find 1 w−(m) 1(cid:88) + 1(cid:88) w+(m)w+(m + ∆m) . . . w+(m(cid:48)) w−(m)w−(m + ∆m) . . . w−(m(cid:48) + ∆m) m(cid:48)=m w+(m)w+(m + ∆m) . . . w+(m(cid:48) − ∆m) w−(m)w−(m + ∆m) . . . w−(m(cid:48)) = m(cid:48)=m (36) where the sum is over all the discrete values of the occu- pancy from m to 1 in increments of ∆m. The compact form in the second line is obtained by understanding the m(cid:48) = m term to have a numerator equal to 1. We can rewrite the terms of this series in a more trans- parent form using Equation (24) above, which relates ra- tios of rates to ratios of steady-state probabilities: w+(m) w−(m + ∆m) = pss(m + ∆m) pss(m) . (37) Substituting this expression into Equation (36), we ob- tain: 1(cid:88) 1(cid:88) m(cid:48)=m ∆τm = = pss(m(cid:48)) m(cid:48)=m w−(m)pss(m) pss(m + ∆m)pss(m + 2∆m) . . . pss(m(cid:48)) w−(m)pss(m)pss(m + ∆m) . . . pss(m(cid:48) − ∆m) (38) where we have canceled out all terms that occur in both the numerator and denominator. Note that the m(cid:48) = m term behaves correctly, since pss(m(cid:48)) will cancel the pss(m) from the denominator, leaving the expected value 1/w−(m). The final step to obtain τm∗,m0 is to note the trivial fact that τm0,m0 = 0, since it takes no time to reach m0 starting from m0. This lets us construct τm∗,m0 by adding up ∆τm's from m0 to m∗: m ∗(cid:88) ∗(cid:88) m m=m0 τdiss ≡ τm∗,m0 = = ∆τm 1(cid:88) pss(m(cid:48)) m=m0 m(cid:48)=m w−(m)pss(m) . (39) This is the expression used in Equation (10) of the main text. In the large N limit, we can recast this in a more in- tuitive form by replacing pss(m) with the asymptotic ex- pression e−N f (m) (noting that the normalization constant cancels out). This yields (cid:90) m ∗ (cid:90) 1 τdiss ≈ N 2 dm m0 m (cid:48) dm(cid:48) eN [f (m)−f (m )] 1 w−(m) . (40) As N → ∞, this integral can be evaluated via Laplace's method by finding the point where f (m)− f (m(cid:48)) is max- imized. This occurs when m(cid:48) is the location of the high- m local minimum m∗, and m is the location of the lo- cal maximum between the two states m†, yielding the asymptotic expansion (cf. [24, eq. 6.4.35]): † τdiss ≈ eN [f (m ∗ )] )−f (m (cid:112) 2πN 1 f(cid:48)(cid:48)(m∗)f(cid:48)(cid:48)(m†) w−(m†) . (41) For large N , the exponential part dominates the qualita- tive behavior of τdiss as the free energy landscape changes, which is the meaning of Equation (11) in the main text. D. Comparison with Equilibrium Statistics Figure 3 of the main text illustrates how local steady- state fluctuations near the high-m local minimum can be approximated by equilibrium distributions with altered parameters. The example at low ∆µ was well approx- imated by the equilibrium free energy landscape at the same temperature and coupling, but reduced monomer concentration cA. For high ∆µ, the matching land- scape had infinite temperature, with constant off-rate per 9 monomer equal to kI = 0.01. Figure 4 confirms the va- lidity of these approximations for the whole range of ∆µ values displayed in Figure 2a, and for both of the βJ val- ues from that figure that exhibit a nonequilibrium phase transition. The low-∆µ approximation works because the local minimum m∗ is only slightly perturbed from its equi- librium value by the active regeneration, and remains within the range of possible values for the high-m lo- cal minimum at the original temperature. In the vicin- ity of m∗, the nonequilibrium contribution fneq(m) ≡ f (m) − feq(m) to the effective free energy landscape can be approximated as a linear function of m: fneq(m) = fneq(m∗) + f(cid:48)neq(m∗)(m− m∗) +O[(m− m∗)2]. Up to an arbitrary additive constant, the effect of the driving is to add an extra linear term mf(cid:48)neq(m∗) to feq. If we look at the form of feq(m) in Equation (7), we see that this is equivalent to changing the monomer concentration to A = cAe−f ceff This argument fails when ∆µ crosses the threshold of the nonequilibrium phase transition, and m∗ moves out- side of the range of values achievable by changing con- centration. In this high-∆µ regime, setting the concen- ) destroys the high-m tration equal to ceff local minimum in the equilibrium landscape, leaving just a single minimum near m = 0. A = cAe−f (cid:48) neq(m (cid:48) neq(m ∗ ∗ ). To understand the fluctuations in this regime, we ob- serve that when kA = kI e−β(∆µ0−Jm) (cid:28) 1, almost every transition to the inactive state ends in dissociation from the lattice, so the dissociation rate through this pathway is equal to the inactivation rate kI . If it is also the case that kI (cid:29) e−βJm, then this is the dominant pathway for dissociation, and we can write w−(m) ≈ N mkI . These same requirements guarantee that w+(m) ≈ N (1 − m)cA (42) (43) as long as cI < cA. The m-dependence of both of these rates is identical to that of the undriven dynam- ics in the absence of inter-particle coupling (βJ = 0), but with the dissociation rate reduced from 1 to kI . The arguments of Section VI B above then yield f (m) ≈ m ln m + (1 − m) ln(1 − m) − m ln cA/kI . If this regime includes the high-m local minimum of f (m), as in Figure 3b, then the steady-state fluctuations will resemble those of the equilibrium system at βJ = 0, and the effective temperature is infinite. 10 FIG. 4. Equilibrium approximations are generalizable. Color online. (a) Effective free energy of the driven system and the equilibrium free energy under a reduced monomer concentration, for ∆µ up to 10 kBT and βJ = 12. (b) Same comparison, but with βJ = 14. (c) Effective free energy of the driven system and the equilibrium free energy at infinite temperature, for ∆µ above 11 kBT and βJ = 12. (d) Same comparison, but with βJ = 14. (a)(b)(c)(d)
1808.04389
2
1808
2018-09-19T09:09:20
A Unique Analytical Solution of the White Matter Standard Model using Linear and Planar Encodings
[ "physics.bio-ph", "q-bio.NC" ]
Diffusion-weighted magnetic resonance imaging in brain white matter probes tissue microstructure and allows for the estimation of compartmental diffusion parameters. Recently, it became apparent that traditional single-direction diffusion encodings are not fully sufficient to resolve the white matter compartmental diffusivities. Multiple diffusion encodings have been suggested to make the problem less ambiguous, however, it still remained unclear whether such protocols would completely solve the problem. Here, we constructively prove that a combination of linear and planar diffusion encodings is enough to determine the parameters of the three compartment white matter model.
physics.bio-ph
physics
A Unique Analytical Solution of the White Matter Standard Model using Linear and Planar Encodings Marco Reisert1,2,3, Valerij G. Kiselev1,3, and Bibek Dhital1,3,4 1Medical Center, Faculty of Medicine, University Freiburg, Germany 2Department of Functional and Stereotactic Neurosurgery, Freiburg 3Department of Medical Physics, Freiburg 4Department of Neurophysics, Max Planck Institute for Human Cognition and Brain Sciences, Leipzig, Germany September 20, 2018 Abstract Diffusion-weighted magnetic resonance imaging in brain white matter probes tissue mi- crostructure and allows for the estimation of compartmental diffusion parameters. Recently, it became apparent that traditional single-direction diffusion encodings are not fully sufficient to resolve the white matter compartmental diffusivities. Multiple diffusion encodings have been suggested to make the problem less ambiguous, however, it still remained unclear whether such protocols would completely solve the problem. Here, we constructively prove that a combi- nation of linear and planar diffusion encodings is enough to determine the parameters of the three compartment white matter model. 1 Introduction For a long time, attempts to multi-compartment modeling in brain white matter (WM) with simple single diffusion encodings [Fieremans et al., 2011, Zhang et al., 2012, Novikov et al., 2018, Reisert et al., 2017] led to ambiguous results [Jelescu et al., 2016, Novikov et al., 2018]. For exam- ple, it was argued in [Fieremans et al., 2011] that intra axonal diffusion is substantially smaller than extra axonal diffusion along the axons, while others argued for the opposite [Zhang et al., 2012, Dhital et al., 2017]. Multiple diffusion encodings offer substantially more information than ordi- nary single diffusion encoding schemes [Jespersen et al., 2013, Westin et al., 2014]. However, most efforts in understanding the additional information gained by such methods were focused on dis- persed single-compartment systems thus revealing apparent measures like eccentricity, microscopic and fractional anisotropy [Jespersen et al., 2013, Westin et al., 2014, Szczepankiewicz et al., 2015]. Recent studies have investigated the benefits of using multiple diffusions encodings to resolve white matter compartmental parameters [Lampinen et al., 2017, Dhital et al., 2018]. For exam- ple, spherical diffusion encodings [Dhital et al., 2017] shows very low kurtosis in white and gray matter, which gives rise to the assumption that traces of the tissue compartments are similar. In [Fieremans et al., 2018] an additional spherical encodings were used to stabilize fits and release constraints. Or, in [Coelho et al., 2017, Reisert et al., 2018] a combination of linear and planar encodings was used with the same intention. Thus, the question arises, what kind of protocol is sufficient to solve the problem uniquely? This short note contributes to the answer of this question. We will show that a combination of linear and planar encodings is indeed enough to provide a unique solution of the full 3-compartment model of brain white matter using O(b2) measurements. The key ingredient of the approach is that a combination of linear and planar measurements provide a direct estimate of the mesoscopic orientation dispersion, without relying on any other concurrent estimates. We further discuss an inherent model property, that, under special conditions, this solution still shows an ambiguity. Finally, we demonstrate by a few counterexamples the inadequacy of linear and spherical encoding to resolve the problem ( taking only O(b2) coefficients). 1 Submitted to Magnetic Resonance in Medicine (MRM) 2 2 The White Matter Model We follow the standard tissue model as proposed in [Novikov et al., 2018, Reisert et al., 2017]. In contrast to [Zhang et al., 2012], in this model both intra and extra-axonal compartments undergo the same convolution with the mesoscopic orientation distribution. For a general encoding matrix B the signal for this model looks as follows S(B) = = d2n M (n, B)f (n) (cid:16) d2n vie− tr(BDn i ) + vee− tr(BDn e ) + vfe− tr(B)Df(cid:17) (1) (2) f (n) (cid:90) (cid:90) n∈S2 n∈S2 where f (n) is an arbitrary, normalized orientation distribution function and M (n, B) the axially symmetric, multi-exponential microstructural model with symmetry axis n. The diffusion tensor of intra- and extra-axonal fractions are parametrized as Dn i = nnT Di, Dn e = nnT ∆e + I3De We now focus on linear encoding Blin = bqqT and planar encoding Bpla = b(I3−qqT )/2, where q is the diffusion gradient direction of modulus one and the b-value b is defined as the trace of the b-matrix. Rewriting the microstructural model in terms of the cosine t = qT n between encoding direction and axon orientation gives, Mlin(t, b) = vie−bDit2 Mpla(t, b) = vie−bDi(1−t2)/2 + vee−b∆e(1−t2)/2−bDe + vfe−bDf + vee−b∆et2−bDe + vfe−bDf (3) (4) In this formulation, the convolution with the mesostructural orientation distribution f (n) takes the form Sα(q, b) = d2n f (n) Mα(qT n, b) (5) (cid:90) n∈S2 where α = linear or α = planar depending on the gradient waveform. Note that Sα(q, b) is normalized in the sense Sα(q, 0) = 1. The key to decouple micro and mesostructural contribution is to work in the domain of spherical harmonics. The spherical convolution turns out to be a product of the two spherical harmonic representations, fl,m and M l α(b), of f (n) and Mα(t, b), respectively. Sl,m α (b) = fl,m M l α(b) . (6) We used here in semi-Schmidt normalization1 as in [Reisert et al., 2017]. The signal is charac- terized by a set of quantities that are rotationally invariant for any signal-generating tissue. Sl α(b) = Sl,m α (b)2 = fl M l α(b) (7) Here fl =(cid:112)(cid:80) m fl,m2 > 0 is the rotation invariant mesoscopic dispersion. For both linear and planar encodings, we define the moments m=−l (cid:118)(cid:117)(cid:117)(cid:116) l(cid:88) (cid:12)(cid:12)(cid:12)(cid:12)b=0 l (n)2 = 1 and (cid:82) d2n Y m (cid:90) 1 dk dbk α(0)) Pl(t) dt 2 Sl α(b) −1 W l,k α := 1 4π = fl sgn(M l (cid:12)(cid:12)(cid:12)(cid:12)b=0 dk dbk Mα(t, b) , (8) (9) 1 In this normalization (cid:80)l l = 0, Mα(qT n, b) =(cid:80) l 2l+1 4π M l α(b)Pl(qT n). l l (n) = Pl(cos θ), where Y m are the spherical harmonics and Pl the Legendre polynomials and θ the polar angle of n. The axial symmetry implies that the spherical harmonics expansion of Mα(qT n, b) contains only components with 2l+1 δl,l(cid:48) δm,m(cid:48) and Y 0 l (n)Y m(cid:48) l(cid:48) (n)∗ = 4π m=−l Y m Submitted to Magnetic Resonance in Medicine (MRM) 3 where sgn(x) = x/x and we do not write the delta-functional term for l ≥ 2, since M l α(b) have definite signs. This follows from their physical meaning of the signal from the idealized unidirectional fiber bundle, since diffusion is faster along such a bundle, M 2 pla > 0, for all meaningful constellations of microstructural parameters. Introduction of these definite signs is sufficient to resolve the ambiguity borne by taking the square of Eq. (7), which is necessary to build rotation invariant quantities. lin < 0 and M 2 Note that the moments defined in Eq. (8) generalize the moments used by [Novikov et al., 2018]; lin ∝ M (2k),l following definitions in [Novikov et al., 2018] equations (15-18). for linear encoding W l,k 2.1 Finding the solution The white matter model described above includes one known (the free water diffusivity, Df) and five unknown scalar parameters: intra-axonal difusivity (Di), extra-axonal radial diffusivity (De), difference between extra-axonal parallel and radial diffusivity (∆e), and the volume fractions (vi, ve, vf) with the constraint vi + ve + vf = 1. The orientation distribution function f (n) contains an infinite set of coefficients. In this section we show that resolving the signal for both linear and planar encoding up to the order b2 and l = 2 enables unambiguous determination of the scalar parameters and the first non-trivial coefficient, f2, of f (n). The corresponding moments are expressed via the model parameters as follows: ∆eve − Deve − 1 3 Divi − Dfvf f2[∆eve + Divi] eve + D2 eve + D2 f vf + lin = lin = − 1 W 0,1 3 2 W 2,1 15 1 5 W 0,2 ∆2 lin = lin = −f2 W 2,2 2 W 0,2 ∆2 pla = 15 pla = −f2 W 2,2 (cid:20) 4 (cid:20) 4 35 1 5 4 35 i vi + D2 4 15 i vi + D2 ∆2 eve + ∆eDeve eve + D2 eve + ∆2 eve + 105 f vf + 2 15 4 105 D2 D2 i vi + D2 2 15 i vi + ∆eDeve 2 3 ∆eDeve (cid:21) 2 3 ∆eDeve (cid:21) (10) (11) (12) (13) (14) (15) The calculations straightforwardly follow from, Eq. (8). Note the absence of the moments W l,1 pla - in this order (linear in b) measurements with any shape of B is equivalent to a set of single-direction measurements and thus do not add any extra information. For example, the signal obtained using the planar encoding in the x, y plane is equivalent to the mean of signals encoded linearly in the x and y directions. In particular, W 0,1 pla , which can be observed from the fact that Mpla(t, b), Eq. (4), can be obtained from Mlin(t, b), Eq. (3), by substituting t2 with (1 − t2)/2 = [P0(t) − P2(t)]/3. Complimentary information can be found in the second (or higher) order of b. pla and W 2,1 lin = 2W 2,1 lin = W 0,1 The dispersion parameter, f2, can be easily found from Eq. (12)-Eq. (15) f2 = − 7 4 lin − 2W 2,2 W 2,2 lin − W 0,2 W 0,2 pla pla (16) Note that the denominator is just the average eccentricity of the compartments [Jespersen et al., 2013] (or microstructural fractional anisotropy), namely W 0,2 lin − W 0,2 pla = ∆2 eve + D2 i vi. Finding other parameters is not so straightforward. Assuming f2 is known, we define a set of Submitted to Magnetic Resonance in Medicine (MRM) auxiliary variables xi as follows pla x1 = 15 W 2,1 lin 2f2 lin − W 0,2 x2 = W 0,2 lin − 5 x3 = −W 0,1 2f2 1 W 2,2 4f2 lin − 45 W 2,2 2f2 x4 = W 0,2 15 2f2 lin + x5 = W 2,1 lin lin + W 2,2 pla = ∆eve + Divi = ∆2 eve + D2 i vi = Deve + Dfvf 9 2f2 W 2,2 pla = D2 eve + D2 f vf = ∆eDeve This system including the constraint on the compartment water fractions vi + ve + vf = 1 defines all scalar parameters. In the following derivation all parameters are restricted to be strictly positive. Let's express all unknowns in terms of vf. From simple algebra applied to Eqs. (19,20) and then Eq. (21) we find De = x4 − D2 f vf x3 − Dfvf , ve = (x3 − Dfvf)2 x4 − D2 f vf , ∆e = x5 x3 − Dfvf The intra-axonal parameters are expressed from Eqs. (17,18), 5 − D2 x2x4 − x2 Di = x2 − ∆2 eve x1 − ∆eve = f vfx2 x1x4 − x3x5 + Dfvfx5 − D2 f vfx1 and vi = (x1 − ∆eve)2 x2 − ∆2 eve = (x1x4 − x3x5 + Dfvfx5 − D2 (x4 − D2 5 − x2x4 + D2 f vf)(x2 f vfx1)2 f vfx2) 4 (17) (18) (19) (20) (21) (22) (23) (24) (25) (26) (27) Now, all expressions depend exclusively on the unknown vf. The last equation vi + ve + vf = 1 solves for vf as follows 1 − vf = vi + ve f vfx2 D2 = 1 − x2 1x4 − x2x2 f x2 + 2x1x3x5 + 2Dfvfx2x3 − 2Dfvfx1x5 3 − D2 f v2 5 − x2x4 + D2 x2 5−x2x4 +D2 f vfx2) (which is allowed since x2 f vfx2 5−x2x4 + Multiplying both sides by the denominator (x2 f vfx2 = −D2 D2 f x2 (x2D2 i D2 1 + 2Dfx1x5− 2x2x3Df− x2 f − D2 evevi (cid:54)= 0) leads to the following equation in vf 5− x2x4) = 0 (28) This equation is linear, since the quadratic terms in vf cancel, which results in the unique final solution 1− 2x1x3x5 + x2x2 5 + x2x4)vf + (x4x2 3 + x2 vf = x2x4 − x2x2 f x2 − x2 3 − x2 5 − D2 1x4 − x2 f x2 5 + 2x1x3x5 1 − 2Dfx2x3 + 2Dfx1x5 x2x4 + D2 (29) By inserting this vf into Eqs. (23 -- 25) we obtain the full solution for all parameters, which is our main result. We now analyze the case of zero denominator in Eq. (29), which will express an ambiguity inherent to the model itself. Substituting the defining expression for the xi's, Eqs. (17 -- 21), gives for the denominator the form vevi(DiDe − DiDf + ∆eDf)2 and the same form multiplied with vf for the numerator. This means that for the special case DiDe − DiDf + ∆eDf = 0 (30) there is no information about vf, since Eq. (28) turns into an identity. In other words, the constraint vi + ve + vf = 1 is automatically fulfilled for any vf. Note that the system of Eqs. (17 -- 21) is linear Submitted to Magnetic Resonance in Medicine (MRM) 5 Figure 1: An example for a family of solutions where DiDe − DiDf + ∆eDf = 0 for varying vf. All these solutions have the same linear and planar moments up to order two. in the volume fractions. In particular, given the diffusivities, Eqs. (17,19,22) can be used to build a linear system for the volume fractions. The determinant of the so constructed system is just the left-hand side of Eq. (30) -- its zero value implies a linear dependency, thus resulting in an infinite number of solutions. To understand the physics behind the degeneracy condition, Eq. (30), consider first two special cases. If De = 0, Eq. (30) gives ∆e = Di, which means that the extra-axonal compartment is indis- tinguishable from the intra-axonal one. Another case is the isotropic extra-axonal compartment, ∆e = 0, in which case it is indistinguishable from free water, De = Df. We found a family of solu- tions that interpolates between these two special cases, which is shown in Figure 1, as a functions of vf. This solution only exists for a specific choice of diffusivities obeying Eq. (30). All the shown solutions have exactly the same moments up to second order. Outside the displayed interval, the solution is unphysical with several negative parameters. Interestingly, the intra-axonal diffusivity is not subjected to the ambiguity. In that case, one can find Di = Dfx2 . Dfx1−x5 Note the similarity of the above degeneracy to the bi-exponential model when the diffusivities in two compartments are equal. Equation (30) expresses this inherent drawback of multi-exponential models exemplified by the standard white matter model. 2.2 Determination of mesoscopic dispersion f2 Equation (16) operates with the moments of the order b2. Here we show that the dispersion can also be expressed directly in terms of the signal. Recall that the moments W l,k α define the Taylor expansion of Sl α(b) in powers of b according to Eq. (8). Therefore the function lin(b) − 2S2 S2 lin(b) − S0 S0 reproduces Eq. (16) with account for the identities W 0,1 one has to consider the function F (b) = f2 + O(b) F (b) := − 7 4 pla(b) pla(b) lin = W 0,1 pla and W 2,1 lin = 2W 2,1 pla . Practically, and fit it linearly to find its value for b = 0. 2.3 Linear and spherical encodings are not sufficient For spherical encoding we have Ssph(b) = vie−bDi/3 + vee−b(∆e/3+De) + vfe−Dfb (cid:12)(cid:12)(cid:12)(cid:12)b=0 W k sph = dk dbk Ssph(b) (31) (32) (33) 00.050.10.150.20.250.30.3500.511.522.5300.050.10.150.20.250.30.3500.10.20.30.40.50.6 Submitted to Magnetic Resonance in Medicine (MRM) 6 Figure 2: Signal courses for the four examples, where linear and spherical moments are identical up to order two. First notable differences appear above b = 2. Note that for S2 lin differences are enlarged by a factor of ten. We assume that only moments up to O(b2) are observable, i.e. W 0,1 lin , W 2,1 sph, W 2 sph are known. We know that W 1 lin , so linear and spherical encodings give five equations up to order 2. In fact, with these equations, one can find analytically a solution for the two-compartment model without the fast water fraction. However, this solution has two roots and is, hence, ambigious. We do not show here the solutions, but give a few numeric examples, where both roots lead to physical meaningful results: sph is linearly dependent on W 0,1 lin , W 0,2 lin and W 0,2 lin , W 2,2 lin , W 1 solution Di 2.00 1.a 2.11 1.b 2.00 2.a 2.17 2.b 3.a 2.40 2.58 3.b 2.00 4.a 4.b 2.14 ∆e 0.60 1.29 0.60 1.11 1.00 1.51 0.60 1.20 De 0.50 0.24 0.50 0.31 0.50 0.31 0.50 0.27 f2 0.80 0.74 0.80 0.72 0.80 0.75 0.50 0.46 vi 0.60 0.31 0.40 0.17 0.40 0.14 0.50 0.24 where mainly vi, ∆e and De are confused. The parameters Di, f2 and ∆e + De are rather stable. This goes in line with the observation that a spherical encoding can resolve the ambiguity of the parallel diffusivities [Fieremans et al., 2011, Fieremans et al., 2018, Dhital et al., 2017], but still has to struggle with ∆e, De and vi. In Figure 2 we show signal courses for the counterexamples. 3 Conclusion We have constructively shown that linear and planar diffusion encodings can fully resolve the three-compartment model of white matter using data up to the order O(b2) and l = 2. The com- mon experience with the diffusional kurtosis imaging [Jensen et al., 2005] indicates the practical availability of O(b2) terms. While in principle, these terms include information for l ≤ 4, the order l = 4 is spoiled by noise as it was shown for a typical two-shell measurement on an advanced scanner with the maximal gradient strength 80 mT/m [Reisert et al., 2017, Fig. 2]. Our analysis highlighted a special situation of ambiguous solution due to an inherent inability of multiexponential models to resolve compartments with similar parameters. The only way to distinguish such compartments is measuring in a domain where their differences get apparent, for example in the large b-value regime, where stable estimates of higher order information becomes possible. Without such information, a stable parameter estimate is only possible relying on prior knowledge. We have also shown that a combination of spherical and linear encoding is not enough to find a unique solution in order O(b2). In fact, O(b2) information delivered by a spherical encoding is fully contained in the combination of linear and planar information, namely W 0,2 lin )/3, which renders a spherical encoding in the presence of linear and planar encodings in the low b- value regime superfluous. In fact, it is a fortunate coincidence that O(b2) information spanned by pla − W 0,2 sph = (4W 0,2 012300.20.40.60.81012300.20.40.60.81012300.20.40.60.81012300.20.40.60.81b-valueb-valueb-valueb-valueexample 1example 2example 3example 4S(b)-10 S2linS0linSspha)b)a)b)a)b) Submitted to Magnetic Resonance in Medicine (MRM) 7 linear and planar diffusion encoding (it is actually the 'full' encoding in O(b2)) is 6 dimensional (Eqs. (10 -- 15)) and the parameter space of the three compartment white matter model has also 6 free parameters, Eq. (2). The derived mapping is only valid for noiseless signals, i.e., when the signal is in the image of the modeling equation. For practical applications the obtainable signal-to-noise ratios are too low. A recent preprint [Coelho et al., 2018] shows by numerical simulations that in a slightly simplified setting (two-compartments and Watson distribution) also in the noisy case the degeneracy is re- solved. The importance of the analytical solution lies in its justification for parameter estimators that rely on unimodal posterior distributions. Additionally, the solution can give certain hints for the construction of such parameter estimators. In fact, the expression of the parameters are all low-order rational functions of the moments (which are all linear projections of the signal). This suggests to make a similar approach for the estimator (e.g. as found in [Reisert et al., 2017]), i.e. using functions, which are rational in linear combinations of the signal. References [Coelho et al., 2017] Coelho, S., Beltrachini, L., Pozo, J., and Frangi, A. (2017). Double diffusion encoding vs single diffusion encoding in parameter estimation of biophysical models in diffusion- weighted mri. In Proceedings of the ISMRM, Honolulu. [Coelho et al., 2018] Coelho, S., Pozo, J. M., Jespersen, S. N., Jones, D. K., and Frangi, A. F. (2018). Double diffusion encoding prevents degeneracy in parameter estimation of biophysical models in diffusion mri. arXiv preprint arXiv:1809.05059, submitted to MRM. [Dhital et al., 2017] Dhital, B., Kellner, E., Kiselev, V. G., and Reisert, M. (2017). The absence of restricted water pool in brain white matter. Neuroimage. [Dhital et al., 2018] Dhital, B., Reisert, M., Kellner, E., and Kiselev, V. G. (2018). Diffusion weighting with linear and planar encoding solves degeneracy in parameter estimation. In Pro- ceedings of the ISMRM, Paris. [Fieremans et al., 2011] Fieremans, E., Jensen, J. H., and Helpern, J. A. (2011). White matter characterization with diffusional kurtosis imaging. Neuroimage, 58(1):177 -- 188. [Fieremans et al., 2018] Fieremans, E., Veraart, J., Benjamin, A.-A., Filip, S., Nilsson, M., and Novikov, D. (2018). Effect of combining linear with spherical tensor encoding on estimating brain microstructural parameters. In Proceedings of the ISMRM, Paris. [Jelescu et al., 2016] Jelescu, I. O., Veraart, J., Fieremans, E., and Novikov, D. S. (2016). De- generacy in model parameter estimation for multi-compartmental diffusion in neuronal tissue. NMR Biomed, 29(1):33 -- 47. [Jensen et al., 2005] Jensen, J. H., Helpern, J. A., Ramani, A., Lu, H., and Kaczynski, K. (2005). Diffusional kurtosis imaging: The quantification of non-gaussian water diffusion by means of magnetic resonance imaging. Magn. Reson. Med., 53:1432 -- 1440. [Jespersen et al., 2013] Jespersen, S. N., Lundell, H., Sønderby, C. K., and Dyrby, T. B. (2013). Orientationally invariant metrics of apparent compartment eccentricity from double pulsed field gradient diffusion experiments. NMR in Biomedicine, 26(12):1647 -- 1662. [Lampinen et al., 2017] Lampinen, B., Szczepankiewicz, F., Mårtensson, J., van Westen, D., Sund- gren, P. C., and Nilsson, M. (2017). Neurite density imaging versus imaging of microscopic anisotropy in diffusion mri: a model comparison using spherical tensor encoding. Neuroimage, 147:517 -- 531. [Novikov et al., 2018] Novikov, D. S., Veraart, J., Jelescu, I. O., and Fieremans, E. (2018). Rotationally-invariant mapping of scalar and orientational metrics of neuronal microstructure with diffusion mri. NeuroImage, 174:518 -- 538. [Reisert et al., 2017] Reisert, M., Kellner, E., Dhital, B., Hennig, J., and Kiselev, V. G. (2017). Disentangling micro from mesostructure by diffusion mri: A bayesian approach. Neuroimage, 147:964 -- 975. Submitted to Magnetic Resonance in Medicine (MRM) 8 [Reisert et al., 2018] Reisert, M., Kiselev, V. G., and Dhital, B. (2018). Unconstrained estima- tion of microstructure by the combination of single- and double-planar diffusion encoding. In Proceedings of the ISMRM, Paris. [Szczepankiewicz et al., 2015] Szczepankiewicz, F., Lasič, S., van Westen, D., Sundgren, P. C., Englund, E., Westin, C.-F., Ståhlberg, F., Lätt, J., Topgaard, D., and Nilsson, M. (2015). Quantification of microscopic diffusion anisotropy disentangles effects of orientation dispersion from microstructure: applications in healthy volunteers and in brain tumors. NeuroImage, 104:241 -- 252. [Westin et al., 2014] Westin, C.-F., Szczepankiewicz, F., Pasternak, O., Özarslan, E., Topgaard, D., Knutsson, H., and Nilsson, M. (2014). Measurement tensors in diffusion mri: generalizing the concept of diffusion encoding. In International Conference on Medical Image Computing and Computer-Assisted Intervention, pages 209 -- 216. Springer. [Zhang et al., 2012] Zhang, H., Schneider, T., Wheeler-Kingshott, C. A., and Alexander, D. C. (2012). Noddi: practical in vivo neurite orientation dispersion and density imaging of the human brain. Neuroimage, 61(4):1000 -- 1016.
1011.2438
1
1011
2010-11-10T17:26:13
Force-extension curves of bacterial flagella
[ "physics.bio-ph", "cond-mat.mes-hall", "cond-mat.soft", "q-bio.CB" ]
Bacterial flagella assume different helical shapes during the tumbling phase of a bacterium but also in response to varying environmental conditions. Force-extension measurements by Darnton and Berg explicitly demonstrate a transformation from the coiled to the normal helical state [N.C. Darnton and H.C. Berg, Biophys. J. {92}, 2230 (2007)]. We here develop an elastic model for the flagellum based on Kirchhoff's theory of an elastic rod that describes such a polymorphic transformation and use resistive force theory to couple the flagellum to the aqueous environment. We present Brownian dynamics simulations that quantitatively reproduce the force-extension curves and study how the ratio $\Gamma$ of torsional to bending rigidity and the extensional rate influence the response of the flagellum. An upper bound for $\Gamma$ is given. Using clamped flagella, we show in an adiabatic approximation that the mean extension, where a local coiled-to-normal transition occurs first, depends on the logarithm of the extensional rate.
physics.bio-ph
physics
EPJ manuscript No. (will be inserted by the editor) 0 1 0 2 v o N 0 1 ] h p - o i b . s c i s y h p [ 1 v 8 3 4 2 . 1 1 0 1 : v i X r a Force-extension curves of bacterial flagella Reinhard Vogel1 and Holger Stark1 1 Institute for Theoretical Physics , TU Berlin 2 Institute for Theoretical Physics , TU Berlin Received: date / Revised version: date Abstract. Bacterial flagella assume different helical shapes during the tumbling phase of a bacterium but also in response to varying environmental conditions. Force-extension measurements by Darnton and Berg explicitly demonstrate a transformation from the coiled to the normal helical state [N. C. Darnton and H. C. Berg, Biophys. J. 92, 2230 (2007)]. We here develop an elastic model for the flagellum based on Kirchhoff's theory of an elastic rod that describes such a polymorphic transformation and use resistive force theory to couple the flagellum to the aqueous environment. We present Brownian dynamics simulations that quantitatively reproduce the force-extension curves and study how the ratio Γ of torsional to bending rigidity and the extensional rate influence the response of the flagellum. An upper bound for Γ is given. Using clamped flagella, we show in an adiabatic approximation that the mean extension, where a local coiled-to-normal transition occurs first, depends on the logarithm of the extensional rate. PACS. PACS-key discribing text of that key -- PACS-key discribing text of that key 1 Introduction Many types of bacteria, such as Escherichia coli and Sal- monella typhimurium, propel themselves forward by rotat- ing a bundle of elastic filaments with helical shape. Each of these flagella, as they are called, is driven by a rotary mo- tor. When the sense of rotation of one motor is reversed, the attached flagellum leaves the bundle and undergoes a sequence of different helical configurations characterized by their pitch, radius and helicity. During the flagellar polymorphism, the bacterium tumbles and then contin- ues swimming in a different direction, when the flagel- lum assumes its original form and returns into the bundle. Whenever the bacterium senses a positive food gradient, it prolongs the swimming phase and thereby increases the food uptake [1]. In recent years, bacteria and their flagella have also been used in studies relevant to microfluidics to transport colloids [2,3] and pump fluids [4] but also to create new liquid-crystalline phases of screw like objects [5]. An understanding of the bacterial tumbling motion and the applications of bacterial flagella in nanotechnol- ogy and microfluidics requires a sufficiently simple elastic model that includes the flagellar polymorphism. This ar- ticle aims to provide such a model. The bacterial flagellum consists of three parts. The ro- tary motor is embedded in the cell membrane and trans- mits its motor torque to the long helical filament with the help of the short and very flexible proximal hook. The filament of E.coli bacteria is up to 10 µm long and is about 0.02 µm in diameter. It is relatively stiff but can switch between distinct polymorphic forms. The filament assumes these forms in response to external perturbations such as changes in pH value, salinity, and temperature of the surrounding fluid [6 -- 8], the addition of alcohols [9] or sugars [10], and by applying external forces or torques to the filament [11 -- 14]. The helical filament is a cylinder formed by 11 protofil- aments each of which consists of a stack of protein mono- mers called flagellin. These monomers assume two con- formations (L and R) that mainly differ in length. Each protofilament only contains one type of monomer and there- fore L- and R-state protofilaments have different lengths.1 Mixing them in the flagellar filament produces bending which together with an intrinsic twist of the filament gives a helical configuration. In total two straight and 10 heli- cal polymorphic states are possible since protofilaments of one type cluster (see Fig. 1). Most of them were ob- served experimentally [15]. This is the picture Asakura [16] and later Calladine [17], with a more detailed geo- metric model, developed to explain the flagellar polymor- phism. The molecular structure for the flagellar filament was confirmed recently [18,19] and several extensions of Calladine's model exist [15,20 -- 23]. This article is mainly motivated by recent experiments of Darnton and Berg [14]. With the help of an optical tweezer set up they pulled the two ends of the flagellar fil- ament apart with a constant velocity and induced a tran- 1 This is one important ingredient of the model developed by Asakura and Calledine. The different helical configurations following from this model were confirmed later for example in Ref. [18]. 2 Reinhard Vogel, Holger Stark: Force-extension curves of bacterial flagella s t r a i g h t n o r m a l h y p e r e x t e n d e d micoile d c u rl y II u rl y I c r a i g h t t s kn e3 f e1 e2 r(s) coiled se n = R 0 1 2 3 4 65 7 8 9 10 11 left-handed right-handed Fig. 1. Twelve polymorphic states of the bacterial flagellum with curvature κ = κmax sin(πnR/11) and torsion τ = τL + (τR − τL)nR/11 after Calladine [17], where nR is the number of protofilaments in the R state. The quantities κmax ≈ 2.4/µm, τL ≈ −5.2/µm, and τR ≈ 11.8/µm are fit parameters. sition between two polymorphic configurations. They then reversed the velocity to compress the flagellum in order to return it to the initial configuration. Darnton and Berg recorded force-extension curves mainly for the transfor- mation from the coiled to the normal configuration. The transformation starts locally at one end of the flagellum and then proceeds in discrete steps along the flagellum. Signatures of the steps are clearly visible in the force- extension curves. Calculating the force-extension curves on the basis of coarse-grained molecular dynamics simulations is not pos- sible since simulation times are far below the experimen- tally relevant time scale of one second [24]. Modeling the polymorphism of a bacterial flagellum on a mesoscopic level is more appropriate. Goldstein et al. extended Kirch- hoff's classic theory of an elastic rod by introducing a double-well potential for the spontaneous torsion to de- scribe the transition between two helical states [25,26]. Wada and Netz also described the helical filament by Kirch- hoff's rod theory but attached a spin variable along the filament in order to distinguish locally between the two helical states [27]. They then performed hybrid Brownian- dynamics Monte-Carlo simulations to numerically calcu- late force-extension curves of bacterial flagella. In this article we present Brownian-dynamics simula- tions of the force-extension curves based on a model that is less time-consuming than the approach of Ref. [27] but that is completely equivalent as we demonstrate in an ap- pendix. We furthermore concentrate on different aspects of the force-extension curves, namely how they depend on the ratio of torsional and bending rigidities Γ and on the velocity or extensional rate with which the flagellum is pulled apart. We also give an upper bound for Γ which is partially in contrast to experimental results. The mean ex- tension, at which a coiled-to-normal transition first occurs locally, is a function of the extension rate. We demonstrate how this extension can be inferred from equilibrium prop- erties of a clamped helical filament. Our modeling is in the spirit of Refs. [25,26]. However, we show that a conven- tional double-well potential cannot reproduce the exper- imentally observed force extension curves. We therefore developed an alternative model, where we just "glue" the Fig. 2. The conformation of a slender elastic rod is described by the space curve r(s) of its center line and the material frame {e1, e2, e3}. The vector κn = ∂se3 describes the local curvature of r(s). harmonic elastic energies of the two helical states together. We perform our simulations with realistic parameter val- ues and can directly compare our results to experiments. The article is organized as follows. In Section 2 we present our extension of Kirchhoff's elastic rod theory to include the polymorphism of helical filaments, explain how we perform the Brownian dynamics simulations, and present analytic results for a uniformly stretched conven- tional helical filament. Section 3 discusses the results for the simulated force-extension curves and addresses the clamped filament. We close with a summary and conclu- sions in Section 4. 2 Continuum model for the bacterial flagellum 2.1 Classic theory of an elastic rod The conformation of a slender rod with contour length L is described by the space curve of its center line r(s), parametrized by the arc length s, and a material frame of three orthogonal unit vectors {e1, e2, e3}, attached to each point on the center line. The vector e3 points along the tangent of r(s) and e1, e2 typically correspond to the principal axes of the inertia tensor of the cross section as indicated in Fig. 2.2 Since the ei are unit vectors, the transport of the material frame along the center line is described by the generalized Frenet-Serret equations ∂sei = Ω × ei, (1) where ∂s means derivative with respect to s. So the con- formation of an elastic rod or filament is completely char- acterized by the angular strain vector Ω = (Ω1, Ω2, Ω3). Alternatively, one uses the Frenet frame consisting of the tangent vector t = e3, the curvature vector κn = ∂st, and the binormal b = t× n. The Frenet frame transforms into the material frame when it is rotated about t = e3 by 2 For a circular cross section the eigenvalues of the inertia tensor degenerate and we are free to choose the vectors e1 and e2. the twist angle φ, as indicated in Fig. 2. This relates the vector Ω to the curvature κ and torsion τ of the filament: Ω1 = κ sin φ, Ω2 = κ cos φ, Ω3 = τ + ∂sφ, (2) (a) fpoly [pN] 3 where the components Ωi are given with respect to the local material frame. Kirchhoff's theory expands the elastic free energy F of a deformed elastic rod up to second order in the angular strain Ω, 0 0 2 k k curvature k [1/mm] 1 (b) 0 azimuthal angle j/L [1/mm] p 0.3 3 t 2 t 1 torsion t [1/mm] 0 3 3 fpoly [pN] 3 0 1 extension z=z/L Reinhard Vogel, Holger Stark: Force-extension curves of bacterial flagella F =Z L 0 fcl(Ω, Ω0)ds fcl(Ω, Ω0) = A 2 (Ω1)2 + A 2 (Ω2 − κ0)2 + C 2 (3) (Ω3 − τ0)2, (4) where we also introduced a spontaneous curvature κ0 and torsion τ0 within Ω0 = (0, κ0, τ0), and A is the bending rigidity and C the torsional rigidity [28,29]. Since the bac- terial flagellum has a circular cross section, we can choose for the material frame the Frenet frame in the undeformed ground state, meaning φ = 0, and thereby restrict the spontaneous curvature κ0 to Ω2. If κ0 and τ0 are con- stant along the filament, it has a helical shape with pitch p = 2πτ0/(κ2 0 ). Note that although free energy (3) is formulated in the spirit of linear elasticity theory, it becomes highly nonlinear when expressed in terms of the space curve r(s) and twist angle φ. 0 ) and radius r = κ0/(κ2 0 + τ 2 0 + τ 2 2.2 Extended Kirchhoff rod theory In order to describe the transition of the bacterial flagel- lum between two polymorphic configurations, the Kirch- hoff rod theory has to be extended to include two local ground states characterized by Ωi = (0, κi, τi) (i = 1, 2). We first tried to generalize the approach of Goldstein et al., who used a typical double well potential for the twist density Ω3 [25,26] (see Appendix A) . However, the force- extension curves calculated with this extended free energy by numerical simulations did not reproduce the experi- mental curves as indicated in Appendix A. We, therefore, developed a different model. To each of the two relevant polymorphic forms of the flagellum we as- sign the elastic free energy (3) of Kirchoff's rod theory and introduce a difference δ of the two energy densities in the ground states. Since the free energy of a system always as- sumes a minimum, we locally assign to the bacterial flagel- lum with angular strain Ω the minimum fpoly(Ω, Ω1, Ω2) of the free energy densities of the two polymorphic con- figurations: fpoly(Ω, Ω1, Ω2) = min(cid:0)fcl(Ω, Ω1), fcl(Ω, Ω2) + δ(cid:1), (5) where fcl is the elastic free energy density of Eq. (4). The resulting density fpoly(Ω, Ω1, Ω2) is plotted in Fig. 3 as a function of curvature κ and torsion τ using Eq. (2) and assuming φ = 0. Fig. 3. (a) Free energy density fpoly from Eq. (5) as a function of curvature κ and torsion τ using Ω = (0, κ, τ ). (b) Free en- ergy density fpoly as a function of height z and azimuthal angle per unit length ϕ/L of a uniformly stretched helical filament (see Section 2.5). The red lines indicate the path of the helical filament in the energy landscape during stretching. Wada and Netz formulated an alternative, statistical model to access the polymorphism of the bacterial flag- ellum [27] which they described as a bead-spring chain. To each bead they assigned the Kirchoff free energy (4) and an Ising spin to distinguish locally between the two polymorphic forms of the flagellum. The Ising spin Hamil- tonian then favored the same polymorphic state for adja- cent beads. By integrating out the spin degree of free- dom, they derived an effective free energy density which we summarize in Appendix B. In experiments, where a thermally induced transition from the normal to the semi- coiled configuration was studied, most flagella assumed a pure polymorphic form of either the normal or semi-coiled state [8] suggesting that the energy cost for forming a do- main wall between the two helical states is much larger than thermal energy. Using this observation, we demon- strate in Appendix B that the free energy density of Wada and Netz simplifies to our elastic free energy (5). Note that our ansatz can easily be extended to describe more than two of the polymorphic states of bacterial flag- ella. In particular, this will be necessary for understanding the sequence of polymorphic transitions during the tum- bling phase of E. coli. The elastic free energy density fpoly of Eq. (5) admits that two domains of different polymorphic states are sep- arated from each other by a sharp domain wall with zero width. To realize a more realistic, smoother transition be- tween two domains, we have to introduce a free energy density of the form [25] fbi(∂sΩ) = γ 2 (∂sΩ)2 , (6) where γ 1/2 is proportional to the width of the domain wall. With a properly adjusted γ, experimental observations of the domain wall could be described quantitatively [25,12]. Furthermore, our simulations demonstrated that the heli- cal filament contains several domains instead of just two when γ is chosen too small or even equal to zero. Instead of implementing a constraint for the filament to be inextensible, we introduce a stretching free energy 4 Reinhard Vogel, Holger Stark: Force-extension curves of bacterial flagella density fst = K (∂sr)2 /2. We choose the spring constant K such that the changes in the filament length are below 1.5 %. So the filament is inextensible to a good approxi- mation. Collecting all the contributions, the total elastic free 0 f ds, which we will use in the following for modeling the bacterial flagellum, is based on the density energy F = R L f =fpoly(Ω, Ω1, Ω2) + fbi(∂sΩ) + fst(∂sr). (7) Since f ultimately depends on the centerline r(s) and the twist angle φ(s), the total free energy is a functional F [r(s), φ(s)] in r(s) and φ(s). 2.3 Dynamics of a Helical Rod We formulate Langevin equations for the location r(s) and intrinsic twist φ(s) of the helical filament. At low Reynolds number the sum of elastic force per unit length, f el = −δF /δr, and thermal force f th is balanced by vis- cous drag. The same applies to the elastic torque per unit length, mel = −δF /δφ and thermal torque mth. Using resistive force theory [30] that employs local friction coef- ficients γk, γ⊥ and γR (see Appendix C), we formulate the Langevin equations (cid:2)γkt ⊗ t + γ⊥(1 − t ⊗ t)(cid:3) v =f el + f th γRω =mel + mth. (8) (9) Here v = ∂tr is the translational velocity, ω = ∂tφ the an- gular velocity about the local tangent vector t = e3, and ⊗ means tensorial product. The anisotropic friction tensor acting on v in Eq. (8) couples rotation about the helical axis to translation and thereby creates the thrust force that pushes the bacterium forward [32]. The friction coef- ficients per unit length were calculated by Lighthill from slender-body theory taking into account the helical geom- etry of the rod [31]. The coefficients are summarized in Appendix C. In experiments one finds reasonable agree- ment with this approach [33,34]. The thermal force f th and torque mth are Gaussian stochastic variables with zero mean, hf thi = 0 and hmthi = 0. Their variances obey the fluctuation-dissipation theorem and therefore read hf th(t, s) ⊗ f th(t′, s′)i = 2kBT δ(t − t′)δ(s − s′) (10) ×hγ−1 t ⊗ t + γ−1 ⊥ (1 − t ⊗ t)i , hmth(t, s)mth(t′, s′)i = 2kBT δ(t − t′)δ(s − s′)γ−1 R , hmth(t, s)f th(t′, s′)i = 0. k (11) (12) 2.4 Details of Simulations In our simulations we used a technique developed by Rei- chert [35] similar to the one of Chirico and Langowski [36] and also employed by Wada and Netz in simulations of helical nano springs etc. [37,38]. We discretize the center- line r(s) of the filament by introducing N + 1 beads at locations ri = r(s = i · h) and with nearest-neighbor dis- tance h. To every bead we attach the material frame, i.e., 2, ei a right-handed tripod of orthonormal vectors {ei 3} (i = 0, . . . , N ), where the tangent vector is approximated by 1, ei e3 i = ri − ri−1 ri − ri−1 . (13) 3 We split the rotation of the tripod along the filament into two parts. First, we rotate about the bond direction e3 i by an angle Ω3 i h corresponding to the intrinsic twist plus tor- sion. Thereafter, the tripod is rotated such that the bond orientation ei 3 is transformed into the consecutive direc- tion ei+1 , thus describing the curvature of the filament. With this procedure the free energy density f of Eq. (7) is discretized and the functional derivatives of the total free energy, f el = −δF /δr and mel = −δF /δφ, reduce to conventional derivatives with respect to ri and φi. We also note that the tangent vector in the friction tensor in Eq. (8) is approximated by ti = (ei . In the following, we will discuss the influence of the three relevant parameters on the force-extension curves; the ratio Γ = C/A of the torsional and bending rigidities (twist-to-bend ratio Γ ), the difference in energy δ of the two helical ground states under study, and the velocity or extensional rate vp with which the bacterial flagellum is pulled apart. The bending rigidity A together with an ap- propriate length introduces characteristic values for force and elastic energy. We used A = 3.5pNµm2 given in Ref. [14] as a typical value for bacterial flagella. All other pa- rameters are determined by the geometry of the two poly- morphic states. Initially, the bacterial flagellum is in the coiled state with spontaneous curvature κ1 = 1.8/µm and torsion τ1 = 0.56/µm and then switches into the normal state with κ2 = 1.3/µm and τ2 = 2.1/µm. 3 + ei+1 3 + ei+1 )/ei 3 3 The friction coefficients are calculated with the formu- las of Lighthill [31] summarized in Appendix C as γk = 1.6 · 10−3pNs/µm2, γ⊥ = 2.8 · 10−3pNs/µm2, and γR = 0.126 · 10−3pNs where a filament diameter of around 20nm was used. The length of the filament is L = 10µm corre- sponding to approximately three helical turns in the coiled and four in the normal state. The discretization length be- tween the beads was chosen as h = 0.2µm and the elastic coefficient in Eq. (6) as γ = 0.1pNµm4. 2.5 Uniform Deformation A helical filament of length L much larger than the pitch deforms uniformly aside from inhomogeneities at both ends when a constant external force pulls at it.3 This means curvature κ and torsion τ are constant along the filament. We orient the helical axis along the z axis of a cylindrical coordinate system (ρ, ϕ, z). The geometry and free energy of the uniformly deformed helix are completely described τ√κ2+τ 2 and the difference in azimuthal by its height z = L 3 For a force compressing the helical filament one observes buckling [40]. Reinhard Vogel, Holger Stark: Force-extension curves of bacterial flagella 5 angles of both ends of the filament, ϕ = ϕ(s = L) = L√κ2 + τ 2, where we set ϕ(s = 0) = 0. Then, the force and torque acting on a uniformly stretched helix follows from F = −∂zF (z, ϕ) and T = −∂ϕF (z, ϕ), respectively. In experiments, the ends of the filament can freely rotate during stretching which means zero torque T . This leads to an expression for ϕ/L ϕ L = κ0 cos α + Γ τ0 sin α (1 + Γ ) cos2 α (14) where we introduced the pitch angle α using the extension ζ = z/L = sin α. (15) Under the condition of zero torque, we then obtain the classic force-extension relation [28] γ⊥ − (γ⊥ − γk)(dz/ds)2. Balancing elastic and viscous forces locally, gives the diffusion equation ∂tz = k γH ∂ 2 s z, (18) where k is the spring constant of the helix given in Eq. (16). A localized deformation at one end of the helical fila- ment spreads diffusively over the whole filament on a time scale τC = γH L2/k ≈ 10−4s which gives the characteristic velocity vC = L/τC ≈ 103 µm/s. In experiments but also in our simulations one pulls at the bacterial flagellum with velocities much smaller than vC . So the extending filament passes through a sequence of equilibrium configurations. This also means that the applied and elastic forces nearly balance each other since they are much larger than the frictional forces. Note that a diffusion equation that bal- ances elastic and viscous forces shows up as well in the twist dynamics of a rotating elastic filament [44]. F (z) = A(Γ κ0 cos α + τ0 sin α)(τ0 cos α − κ0 sin α) (16) cos α(Γ cos2 α + sin2 α)2 z − z0 z0 − z 0 + τ 2 := −k L (κ2 κ2 0(Γ κ2 0 )3 0 + τ 2 0 ) ≈ A L , 3 Force-extension curves In the second line of Eq. (16), the force is linearized in a small relative extension (z − z0)/L, where z0 is the height of the undeformed helix and k the spring constant. Interestingly, we find that curvature κ(ζ) = p1 − ζ 2ϕ/L and torsion τ (ζ) = ζϕ/L of a helix with extension ζ = z/L lie on the ellipse [κ(ζ) − κ0/2]2 + Γ [τ (ζ) − τ0/2]2 = (κ0/2)2 + Γ (τ0/2)2, (17) the center of which is at (κ0/2, τ0/2). In deriving Eq. (17) we used expression (14) for ϕ/L. For a symmetric poten- tial with twist-to-bend ratio Γ = 1, the ellipse becomes a circle. We use relation (17) in Figs. 3(a), 5(b), and 10 to indicate the path of a uniformly stretched flagellum in the κ, τ plane. Bacterial flagella only have a length of a few pitches. So finite size effects from both ends lead to a noticeable dependence of curvature and torsion on arc length s. Nev- ertheless, we find that Eq. (16) is a good approximation for the initial part of the force-extension curve. This is valid since the experimental pulling rate is so small that the extending filament passes through a sequence of equi- librium configurations as we demonstrate in the following section. 2.6 Characteristic Time Scale and Velocity Applying instantaneously a force at one end of the heli- cal filament induces a localized deformation which spreads along the helix. To treat this situation approximately, we assume the helical filament to be an extensible rod that locally obeys the linearized force-extension relation of Eq. (16) with (z − z0)/L replaced by dz/ds. In ad- dition, the rod possesses the local friction coefficient per unit length for moving a helix parallel to its axis, γH = 3.1 Pulling on and compressing the helical filament We performed both Stokesian (temperature T = 0) and Brownian dynamics (T = 300K) simulations of the force- extension measurements in Ref. [14]. Similar to the ex- perimental setup, we fix one end of the filament whereas the other end is allowed to move in a harmonic potential with trap stiffness 100 pN/µm (as in Ref. [14]) mimicking the experimental situation where a bead attached to the bacterial flagellum was trapped by an optical tweezer. The axis of the helical filament is oriented parallel to the z axis. In the beginning, the minimum of the harmonic potential coincides with one end of the filament in the initial coiled state. We then move the potential with a constant veloc- ity vp along the z axis and the filament stretches. After reaching a maximum extension ζM = zM /L, the potential moves with the opposite velocity −vp back into the ini- tial position. Note that in experiments the extension rates were vp = 0.4µm/s and less [14]. Such small velocities are very time consuming in our simulations so that we typi- cally chose values from the range vp = 2, . . . , 20µm/s for recording the complete force-extension curve. When we were just monitoring the initial part of the curve, we used velocities as small as vp = 0.2µm/s (see Section 3.2). We first simulated the extension of a filament with only one helical state. Similar to the snapshots 1 and 2 in Fig. 4(a) the deformation of the filament is uniform except for small regions at both ends. This leads to a small deviation of the simulated force-extension curve from the theoreti- cal prediction of Eq. (16) for a uniformly stretched fila- ment. The situation is similar to the results in Fig. 4(b), where the initial part 1-2 of the simulated blue curve is compared to the analytic prediction (thin black line). If we even shift the thin black line to the right, the dashed black line agrees very well with the blue curve besides the initial part close to position 1. We checked that even this difference gradually vanishes with increasing height of Reinhard Vogel, Holger Stark: Force-extension curves of bacterial flagella (b) 5 time T>0 single runs average T=0 theory 3 vp ] N p [ e c r o f 2 3 4 5 6 coiled v p stretched coiled normal coiled 6 (a) 1 2 3 4 5 6 7 tary movie 2. Clearly, the first transition into the normal state occurs at a smaller extension compared to the deter- ministic case since thermal fluctuations help to overcome the energy barrier in the elastic free energy density (see Fig. 3). Whereas the force in each realization fluctuates visibly, the sharp decrease of the force when a local tran- sition into the normal state occurs is as pronounced as in the deterministic case. Finally, when both ends of the fil- ament are moved together, it completely transforms back to the coiled state and buckling is not observed. The sin- gle realizations of the complete cycle of the force-extension curve closely resemble the experimental curves in Figs. 5 and 6 of Ref. [14]. In particular, the force and extension where the first coiled-to-normal transition occurs fall into the experimental ranges of 3 to 5 pN and ζ = 0.55 to 0.6, respectively. We now discuss in more detail how the difference δ in the ground-state energies of the two helical states, the twist-bend ratio Γ , and the extension rate vp influence the force-extension curves. 3.1.1 Ground-state energy difference δ of the coiled and normal state Increasing the ground-state energy δ > 0 of the normal relative to the coiled state also increases the energy bar- rier, which the flagellum in the coiled configuration has to overcome to transform locally into the normal state. Therefore, the transition is delayed to a larger extension z/L or does not occur at all. On the other hand, the bar- rier which the normal configuration has to overcome to relax back into the coiled state decreases and buckling of the filament becomes less probable. Observations also show that a filament prepared in the normal state very slowly relaxes back into the coiled state, so δ should not be too large compared to thermal energy kBT . Using also the following quantitative considerations, we adjusted δ to 0.1pN which resulted in the good agreement with experi- mental observations, already demonstrated. We now derive an upper bound for δ to ensure that a transformation from the coiled to the normal state is, in principal, observable. The locus of the energy barrier in the κ, τ plane of Fig. 3 is given by fcl(Ω, Ω1) = fcl(Ω, Ω2) + δ, (19) where Ω = (0, κ, τ ) and Ωi = (0, κi, τi) contains the val- ues for spontaneous bend κi and torsion τi of the coiled (i = 1) and normal state (i = 2), respectively. Assum- ing both states have the same elastic constants A, C, we arrive at a straight line κ(κ1 − κ2) + τ Γ (τ1 − τ2) + δ/A = (κ2 1 − κ2 2)/2 + Γ (τ 2 1 − τ 2 2 )/2. (20) We will use this formula later. The ground-state energy density of the normal state at Ω2 = (0, κ2, τ2) is δ. Only if δ lies below the energy density of the coiled state at Ω = Ω2 is a clear transition between both states possible. buckled -v p 0 -1 1 7 0.3 0.5 0.7 (coiled) extension z 0.85 (normal) Fig. 4. (a) Snapshots of a helical filament stretched at T = 0 with velocity vp = 2µm/s taken from the supplementary movie 1. The coiled-to-normal transition (vp > 0) and buckling (vp < 0) are visible. (b) Force-extension curves simulated without thermal noise (T = 0) (blue curve) and with thermal noise: two realizations (thin orange and green lines) and an average over 10 runs (thick red line) are shown. The parameters are vp = 2µm/s and Γ = 0.7. Analytic prediction for a uniformly stretched filament (thin black line) and shifted curve (dashed black line). the helix as expected. In Section 2.6 we already explained that our simulations are performed in the quasistation- ary regime. This is also confirmed by our observation that at an extension rate vp = 2µm/s the dissipated energy during one extension-and-compression cycle is only about 2% of the maximum elastic energy at extension ζM . Fur- thermore, we do not observe any pronounced difference between deterministic Stokesian and Brownian dynamics simulations. For T 6= 0, the forces fluctuate around a value which agrees with the deterministic force at T = 0. We then determined the force-extension curve at T = 0 when the helical filament can switch from the coiled to the normal state using the elastic free energy (7). The re- sults are illustrated in Fig. 4, where the blue curve in (b) corresponds to T = 0. At a certain extension (position 2) the measured force drops sharply since a small seg- ment of the filament close to the fixed end switches into the normal state as snapshots 2 and 3 in Fig. 4(a) reveal. Then the filament is stretched again. Further adjacent seg- ments transform suddenly until at position 6 the filament is nearly completely in the normal state. From here we invert the velocity vp and move both ends together. How- ever, the filament does not transform back into the coiled state but remains in the normal state. This ultimately leads to a negative force under which the filament buckles [see snapshot 7 in Fig. 4(a)]. The full cycle is shown in the supplementary movie 1. Brownian dynamics simulations reveal the influence of thermal fluctuations on the force-extension curve. Two re- alizations are included in Fig. 4(b) as thin lines and the thick red line shows an ensemble average over 10 runs. A full cycle of one realization is shown in the supplemen- Reinhard Vogel, Holger Stark: Force-extension curves of bacterial flagella 7 The explicit form of this upper bound, δ < fcl(Ω2, Ω1), combines with δ > 0 to the inequality 0 < δ A(κ1 − κ2)2 <(1 + Γ ∆2)/2, (21) (22) where ∆ = τ1 − τ2 κ1 − κ2 is the ratio of the differences in spontaneous torsion and curvature. In experiments the value of δ changes with the condi- tions of the solvent. In particular, different polymorphic forms of the bacterial flagellum become stable when one alters the pH value, ionic strength, or temperature of the aqueous environment [8]. It would be interesting to per- form the force-extension experiments under different con- ditions to investigate how a changing δ but also variations in bending (A) and torsional (C) rigidity influence the force-extension curve of a bacterial flagellum. 3.1.2 Twist-to-bend ratio Γ Increasing the twist-to-bend ratio Γ also increases the en- ergy barrier as the following formula for the minimum value of the barrier demonstrates, fb A(κ1 − κ2)2 =(cid:0)1 + 2δ/[A(κ1 − κ2)2] + Γ ∆2(cid:1)2 8 (1 + Γ ∆2) . (23) However, in contrast to δ an increasing Γ increases both barriers for the coiled-to-normal and the normal-to-coiled transition. We now discuss in Fig. 5(a) how the twist-to- bend ratio Γ influences the force-extension curve of the helical filament. The filament is stretched to an extension ζ = 0.95, well above the equilibrium height ζ = 0.85 of the normal state. Fig. 5(b) shows contour plots of the elastic free energy density as a function of curvature and torsion for the same Γ as in (a). The red line indicates the sequence of κ, τ values, as the filament in the coiled state is uniformly stretched (see Section 2.5). Note that for Γ 6= 1 the anisotropy of the elastic free energy density is clearly visible. For Γ = 0.5 and 0.7, one observes the typical force- extension curves as already discussed in Fig. 4 also for Γ = 0.7. Since now the maximal extension ζ = 0.95 is larger, the curve for Γ = 0.7 during compression looks different. While at ζ = 0.95 the entire filament is in the normal state, the filament for Γ = 1 no longer transforms completely into the normal state. Most pronounced at Γ = 1 is the fact that even with thermal forces the filament does not return into the coiled state during compression. This results in a negative force under which the filament buckles. All the force-extension curves in Fig. 5(a) are de- termined for an extension rate vp = 2µm/s. A smaller vp enhances the probability that thermal fluctuations trans- form the filament back into the coiled state and buckling is not observed. The same is true for a smaller extension where a larger part of the filament stays in the coiled state and makes it easier for the rest of the filament to return to the coiled state. At a ratio Γ = 1.5 the force-extension curve changes qualitatively. A first transition into the nor- mal state occurs at a larger extension ζ ≈ 0.65 compared to the previous curves. The transitions are no longer as pronounced and there are larger differences between sin- gle realizations of the force-extension curve. In addition, for Γ < 1.5 only one normal domain occurs whereas for Γ = 1.5 two normal domains are observed. The reason for all these features becomes clear from Fig. 5(b). At Γ = 1.5 a uniformly stretched filament does not hit the energy bar- rier anymore but passes the barrier in a close distance so that thermal fluctuations are needed to induce transfor- mations into the normal state. For further increase of Γ a transition into the normal state does not occur at all realizations, as the three graphs for Γ = 2 demonstrate. A similar behavior is observed in Ref. [14] for a transition from the normal into the hyperextended state. The blue curve corresponds to the traditional force-extension rela- tion of the helix in the coiled state. In the yellow curve a complete transition into the normal state was realized. A partial transformation occurred in the magenta curve which then completely returned into the coiled state. Fi- nally, at Γ = 2.5 the filament always remains in the coiled state. Beyond an extension of ζ = 0.5 the slope of the curve becomes smaller. This is the onset of a qualitatively new behavior. At even larger Γ and τ0/κ0 < 1, one ob- serves a sharp drop in the force due to a discontinuous transformation, where one turn of the helical filament un- winds. This is discussed in Refs. [39,37]. As discussed, neglecting thermal fluctuations, a coiled- to-normal transition is no longer observable in the force- extension curve when in Fig. 5(b) the straight line [Eq. (20)] separating the coiled and normal state becomes tan- gential to the trajectory [Eq. (17)] of the uniformly stretched filament. Combining both equations leads to a second con- dition for δ/A: δ A(κ1 − κ2)2 < 1 2(κ1 − κ2)(cid:18)κ2 + Γ τ2∆ +q(1 + Γ ∆2)(κ2 1 + Γ τ 2 1 )(cid:19) (24) Together with condition (21), we obtain a region in the pa- rameter space (Γ, δ/A) where a transformation from the coiled to the normal state should occur. The region is in- dicated as shaded area in Fig. 6. Based on experimental data for Young's and shear modulus in literature, Flynn and Ma received for the twist-to-bend ratio Γ the range 2 · 10−1−2 · 102 [41]. With a computational method, called the quantized elastic deformational model, they calculated Γ ≈ 23 [41]. On the other hand, together with our value δ/[A(κ1 − κ2)2] ≈ 0.11, we predict a value of Γ . 1, in good agreement with our simulations. 3.1.3 Extensional rate In Section 2.6 we reasoned that during the measurements of the force-extension curves the bacterial flagellum goes 8 Reinhard Vogel, Holger Stark: Force-extension curves of bacterial flagella 0.5 0.7 0.85 1 0.3 0.5 0.7 0.85 0.3 5 G=0.5 G=0.7 ] N p [ e c r o f 0 -1 10 ] N p [ e c r o f 5 0 15 10 5 ] N p [ e c r o f G=1.0 G=1.5 G=2.0 G=2.5 G=0.5 G=0.7 3 t 2 t 1 0 0 k2 k1 3 0 k2 k1 G=1.0 G=1.5 3 t 2 t 1 0 3 t 2 0 k2 k1 3 0 k2 k1 G=2.0 G=2.5 1 5 0 -1 10 5 0 15 10 5 0 1 ] m m / 1 [ t n o i s r o t ] m m / 1 [ t n o i s r o t ] m m / 1 [ t n o i s r o t 3 t 2 t 1 0 3 3 t 2 t 1 0 3 3 t 2 t 1 0 0 0.3 (coiled) 0.5 0.7 extension z 0.85 (normal) 1 0.3 0.5 0.7 (coiled) extension z 0.85 (normal) t 1 0 0 k2 k1 3 curvature k [1/mm] 0 k2 k1 3 curvature k [1/mm] Fig. 5. (a) One or several realizations of force-extension curves for increasing twist-to-bend ratios Γ . The extensional rate is vp = 2µm/s. (b) Contour plots of the elastic free energy density as a function of curvature and torsion for the same Γ as in (a). The red line indicates the sequence of κ, τ values, as the filament in the coiled state is uniformly stretched. 2  2 Κ - 1 Κ  A  ∆ 3 2 1 0 0 0.5 G 1 1.5 Fig. 6. The shaded area indicates the parameter ranges for the ground-state energy difference δ and the twist-to-bend ratio Γ for which a coiled-to-normal transition should be observed. The dots give parameter values used in simulations. Note that thermal fluctuations shift the upper border for Γ indicated by the blue line to the right. (a) 5 3 ] N p [ e c r o f 0 -1 0.3 (coiled) v=20mm v= 5mm v= 2mm G=0.7 (b) 5 ] N p [ e c r o f 3 0 v=20mm v= 5mm v= 2mm G=1 0.5 0.7 extension z 0.85 (normal) -1 0.3 (coiled) 0.5 0.7 extension z 0.85 (normal) Fig. 7. Force-extension curves for twist-to-bend ratios Γ = 0.7 (a) and Γ = 1 (b) at different extension rates vp = 2µm/s, 5µm/s and 20µm/s. through a sequence of equilibrium states. This means that frictional forces acting on the filament through the sol- vent are negligible against elastic forces. Therefore, with- out thermal noise, the force-extension curve does not de- pend on the extensional rate vp. In contrast, our Brown- ian dynamics simulations demonstrate a clear influence of vp. In Fig. 7, we show force-extension curves for twist-to- bend ratios Γ = 0.7 and 1 and different extension rates vp = 2µm/s, 5µm/s and 20µm/s. The first transition from the coiled to normal state occurs at larger extensions when vp is increased. This is immediately clear since smaller ve- locities vp give the filament more time to explore the en- Reinhard Vogel, Holger Stark: Force-extension curves of bacterial flagella 9 ] m m N p [ y g r e n e 4 3 0 t time [s] 1 (a) 75 v =p G= 0.2 mm/s 20 mm/s 0.7 1 ) p v , z ( p 50 25 G = 1 G = 0.7 102 100 ] s [ > t < T P F M -2 10 0.56 0.58 extension z 0.60 Fig. 8. Mean-first-passage time (MFPT) as a function of ex- tension ζ for Γ = 0.7 and Γ = 1.0. Inset: When the local coiled-to-normal transition occurs, the energy decreases instan- taneously. ergy landscape with the help of thermal fluctuations. So the appropriate local curvature and torsion to overcome the potential barrier can be created at smaller extension. The curves in Fig. 7 just give one specific realization for each parameter set. In the following section, we will inves- tigate in detail the probability distribution for the exten- sion where the first coiled-to-normal transition occurs. If the filament in the normal state is compressed too fast, it will start to buckle since again it has not sufficient time to overcome the energy barrier. This is visible in Fig. 7(a), where the filament with the highest compressing rate vp = 20µm/s buckles which corresponds to a negative force. On the other hand, for smaller rates vp, the filament always returns into the coiled state. 3.2 Clamped filament Since the bacterial flagellum goes through a sequence of equilibrium states during the measurements of the force- extension curves, it should be possible to derive some char- acteristics of these curves by investigating a clamped fila- ment which is hold at a fixed extension ζ = z/L in thermal equilibrium. In particular, we show here how one can in- fer the mean extension hζi(vp) at which the filament in a force-extension measurement undergoes the first local transition to the normal state We stretch the filament in the coiled state at zero tem- perature to an extension ζ and keep this extension con- stant. We then perform a Brownian dynamics simulation and determine for each realization the first-passage time τ at which a local transition to the normal state occurs. The inset of Fig. 8 shows the filament before and after the transition accompanied by a sharp decrease of the elastic free energy. According to Kramers theory, the mean-first- passage time (MFPT) is proportional to the Arrhenius factor, hτi ∼ exp(cid:18)△F (ζ) kBT (cid:19) . (25) 0 0.50 (b) 0.62 z n o i s n e t x e 0.58 0.55 0.60 0.65 extension z mean extension <z> simulations G=1 G=0.7 0.54 -1 10 100 101 2 10 extensional rate v [mm/s] p Fig. 9. (a) Probability distributions p(ζ, vp) as a function of extension ζ for Γ = 0.7 (blue) and Γ = 1 (magenta) at exten- sion rates vp = 0.2µm/s (full lines) and vp = 20µm/s (dashed lines). (b) Analytically determined mean extension hζi (vp) for the coiled-to-normal transition as a function of velocity vp for Γ = 0.7 (blue) and Γ = 1 (magenta). The dots indicate nu- merically determined extensions for each simulation run. The activation energy △F (ζ) depends on the extension ζ and is needed to create the domain wall between the coiled and normal state. We determined the MFPT by averaging over 100 simulated values for τ at each extension ζ. The results are plotted in Fig. 8 for two twist-to-bend ratios Γ = 0.7 and Γ = 1. We only simulated hτi for a small variation in the extension for two reasons: at larger ex- tensions where △F (ζ) ≈ kBT the Kramers theory is no longer valid and hτi is too small to be determined accu- rately; at smaller extensions hτi is so large that it cannot be calculated in reasonable simulation times. Note that for Γ = 1 the MFPT spans three decades. Our simulation results demonstrate that log hτi (ζ) can be fitted by log hτi (ζ) ≈ α − βζ, (26) where hτi is measured in seconds. Equation (26) can just be viewed as a Taylor expansion to linear order in ζ. The parameters α and β follow from a least-square fit. Whereas α changes with Γ , the slope β surprisingly does not seem to depend on Γ within the numerical accuracy. A similar law as Eq. (26) is used in situations where bonds rupture under a given load force [43,45]. Note, however, that here we control the extension. Within the adiabatic or quasistationary approxima- tion, we now formulate the probability p(t, vp)dt that with- in the time interval [t, t + dt] the filament transforms lo- cally from the coiled to normal state when stretched with 10 Reinhard Vogel, Holger Stark: Force-extension curves of bacterial flagella exp(cid:18)−Z t 0 1/ hτi (ζ(t′, vp))dt′(cid:19) , (27) (28) velocity vp, p(t, vp) = 1 hτi (ζ(t, vp)) where ζ(t, vp) = ζ0 + tvp/L is the extension at time t. The exponential factor in Eq. (27) is the probability that until time t a transition does not occur, which follows by writing the probability that a transition within the time interval dt′ does not hap- pen as (1 − dt′/ hτi) ≈ exp(−dt′/ hτi). The first factor is the probability per unit time that the transition occurs within dt at time t when the filament is with certainty in the coiled state before. Using Eq. (28), one introduces the probability p(ζ, vp)dζ = p(t, vp)dt that the filament undergoes a local coiled-to-normal transition at extension ζ. We calculate it analytically in Appendix D by assum- ing the validity of Eq. (26) for the whole ζ range. The results are plotted in Fig. 9(a) for Γ = 0.7 (blue) and Γ = 1 (magenta) and for extension rates vp = 0.2µm/s (full lines) and vp = 20µm/s (dashed lines). The probabil- ity distributions are concentrated on a small range about their maximum values and are shifted to larger extensions for increasing velocities vp, as expected. We then calculate the mean extension hζi (vp) = Z ∞ ζ0 ζ′p(ζ′, vp)dζ′, (29) at which the coiled-normal transition occurs first for a given extension rate vp. The evaluation of the integral is performed in Appendix D and finally gives hζi (vp) ∝ log vp. (30) The mean extension hζi (vp) with all prefactors and con- stant terms is plotted in Fig. 9(b) for Γ = 0.7 (blue) and Γ = 1 (magenta) together with numerically determined extensions ζ for several realizations at a specific velocity vp. The values scatter around the mean value and are, therefore, in good agreement with the analytical treat- ment. We note that a relation similar to Eq. (30) occurs for the velocity dependence of the rupture force of single molecular bonds in dynamic force spectroscopy [43]. 4 Summary and conclusions In this article we have developed a sufficiently simple elas- tic model to describe the polymorphism of a bacterial flagellum based on Kirchhoff's theory of an elastic rod. The friction with the aqueous environment is modeled within resistive force theory. Using geometrical parame- ters of the coiled and normal states and the bending rigid- ity as obtained in Ref. [14], we are able to reproduce the force-extension curves recorded in experiments. Thermal fluctuations realized within Brownian dynamics simula- tions are crucial. In particular, the force values at which a first coiled-to-normal transition takes place lie between 3 and 5 pN, as in experiments [14]. We have investigated in detail how the force-extension curve depends on the twist-to-bend ratio Γ and, furthermore, by analytic argu- ments identified a parameter region for ground-state en- ergy difference δ and Γ , where a coiled-to-normal transi- tion should be observable. It clearly demonstrates that for values of Γ well above one, a polymorphic transformation is not possible and therefore contradicts some of the ex- perimental values for Γ recorded in literature. Based on our simulations, we predict Γ to be within 0.7 and 1.0. Further studies demonstrate how the extensional rate vp influences the force extension curve. We directly observe the influence in the simulations of the full force-extension cycle but also when we concentrate on the extension for the first coiled-to-normal transition. Since the extensional rates vp are sufficiently small, the flagellum goes through a sequence of quasi-stationary states. We, therefore, used equilibrium properties of clamped flagella to predict a logarithmic velocity dependence for the mean extension of the first coiled-to-normal transition in good agreement with our simulations. Our approach is easily extended to include more than two polymorphic states. Figure 10 shows the contour lines of the resulting free energy landscape when all twelve poly- morphic conformations of Fig. 1 are taken into account. For each of these conformations we take Kirchhoff's elas- tic free energy with spontaneous curvature and torsion as given by Calladine [17] and just consider the minimum value from all these free energies. We assume here that the ground-state energies of all helical states are zero and that they all have the same bending and torsional rigidi- ties with Γ = 0.7. We also included the paths of uniformly stretched helical filaments. The coiled-to-normal transi- tion (nR = 3 to 2) takes place with certainty. However, all the other transitions would need thermal fluctuations to occur. For example, the normal-to-hyperextended tran- sition (nR = 2 to 1) will only occur when it is stretched sufficiently slowly so that thermal fluctuations can induce the transformation. We can now use our model to study various aspects connected to bacterial locomotion. For example, we will investigate how polymorphic transitions are induced by rotating the flagellum. Our very challenging goal is to model the complete tumbling cycle of a bacterium. Fi- nally, we note that our model may be applicable as well to spirochetes, where very recent experiments and theory have begun to address the elasticity of the helical structure [46]. We thank E. Frey, R. Netz, and A. Vilfan for stimulating dis- cussions and acknowledge financial support from the VW foun- dation within the program "Computational Soft Matter and Biophysics" (grant no. I/83 942). Reinhard Vogel, Holger Stark: Force-extension curves of bacterial flagella 11 left-handed right-handed n = R 0 1 2 3 4 5 6 7 8 9 10 11 e r u t a v r u c ] m m / 1 [ k 3 0 -6 0 torsion t [1/mm] 6 12 Fig. 10. Contour lines of the free energy density of the extended Kirchhoff rod theory that includes all 12 polymorphic states of the bacterial flagellum. For all helical states the same ground state energy and the same bending and torsional rigidity with Γ = 0.7 is assumed. The red lines indicate the paths of uniformly stretched flagella. A Extended Kirchhoff rod theory: the double-well potential Here, the main idea is to extend the theory of Goldstein et al. [25,26]. They used a double well potential for the twist density Ω3, realized by a polynomial of degree four, to describe two polymorphic states of a flagellum [25,26]. In order to develop a strategy how to generalize this double well potential to all three coordinates Ωi (i = 1, 2, 3), we first write down a general one-dimensional poly- nomial of degree four: f (x, x1, x2) = A 2 (x − x1)2 (x1 − x2)2 [(x − x2)2 − d 6 (x − x1)(3x + x1 − 4x2)]. (31) For d < 1 it has two minima at x1 and x2 with f (x1) = 0 and f (x2) = δ, respectively, where δ = 1 12 A(x1 − x2)2d. (32) Whereas ∂ 2 pends on the parameters x1, x2, d, and A. xf (x1) = A, the second derivative at x2 de- We generalize the polynomial of Eq. (31) to three di- mensions by replacing terms of the form Axy by x· Ay, where x, y are three-dimensional vectors and A is a di- agonal matrix with A11 = A22 = A and A33 = C. The constants A, C are the bending and torsional rigidities, respectively. Using the shorthand notation x2 A = x · Ax with x = Ω − Ωi, we now write Ω − Ω12 Ω1 − Ω22 d (Ω − Ω1)· A(3Ω + Ω1 − 4Ω2)] 6 [Ω − Ω22 f (Ω, Ω1, Ω2) = − 1 2 A A A (33) with δ = 1 12Ω1 − Ω22 Ad. (34) The polynomial is illustrated in Fig. 11(a) with Ω = (0, κ, τ ). It has two minima at Ω1, Ω2 with f (Ω1) = 0, (a) f [pN] 0.75 0 0 t 2 torsio t [1/m m ] n (b) 2 ] N p [ e c r o f 1 0 0.3 (coiled) 0.5 extension z 0.7 0.85 (normal) 3 k 1 k 2 curvature k [1/m m ] t 1 0 3 Fig. 11. (a) Double-well potential of Eq. (33) plotted as a function of Ω2 = κ and Ω3 = τ for the parameters of the coiled and normal helical state. (b) Force-extension curve (red line) simulated with this potential at T = 0 and vp = 2µm/s. The blue curve is taken from Fig. 5(a) for the same parameter Γ = 1. f (Ω2) = δ, respectively, and one saddle point at Ω3 = Ω1 + 1 2−d (Ω 2 − Ω1) for d < 1. Close to the first minimum at Ω1, the polynomial agrees with Kirchhoff's elastic free energy (4), whereas the bending and torsional rigidities at Ω2 and also the energy barrier depend on δ. Figure 11(b) shows a force-extension curve (red line) simulated with Eq. (33) at T = 0 for the same parame- ters as in Fig. 5(a) with Γ = 1. For comparison the initial part of the curve from Fig. 5(a) (Γ = 1) is included. The helical filament is much softer and the initial part of the force-extension relation has a negative curvature in con- trast to experiments. We, therefore, decided to introduce the alternative model of Eq. (5). B Simplifying the free energy of Wada and Netz After integrating out the spin degree of freedom, Wada and Netz arrived at the following elastic free energy den- sity [27]: f (Ω, Ω1, Ω2) = f1 − f2 (35) 12 with f1 = A 2 Ω2 1 + A 2 (cid:18)Ω2 − κ1 + κ2 2 (cid:19)2 + C 2 (cid:18)Ω3 − τ1 + τ2 2 (cid:19)2 f2 = kBT a lnh cosh(Λ) +qsinh2(Λ) + e−4J/kB Ti + J a (36) (37) and kBT a Λ =δ + A 2 (κ1 − κ2)(cid:18)Ω2 − κ1 + κ2 2 (cid:19) + C 2 (τ1 − τ2)(cid:18)Ω3 − τ1 + τ2 2 (cid:19) (38) Here A, C are the bending and torsional rigidities, respec- tively, δ is the difference of the ground-state energies of the helical conformations (κ1, τ1) and (κ2, τ2), and a the length of discretization. The quantity J is the interaction strength in the Ising Hamiltonian. We interpret the energy cost 2J for two anti-parallel spins as the energy of a domain wall connecting two he- lical states. Since such domain walls are rarely seen in experiments [8], we can assume 2J ≫ kBT and, therefore, approximate f2 as f2 ≈ kBT a Λ + J a , (39) where we have used cosh x + sinh x = expx. Now, we introduce the elastic free energy densities of the two helical states, α = β = A 2 A 2 Ω2 1 + Ω2 1 + A 2 A 2 (Ω2 − κ2)2 + (Ω2 − κ1)2 + C 2 C 2 (Ω3 − τ2)2 + δ/2, (Ω3 − τ1)2 − δ/2, (40) (41) and write the two contributions to the energy density (35) as f1 = 1 2 (α + β) + c1, f2 ≈ 1 2α − β + J a , (42) where c1 is just a constant. This finally gives our ansatz for the free energy density (5): 1 2 f ≈ (α + β − α − β) + c1 = min(α, β) + c1. (43) C Friction coefficients per unit length In a moving helical filament, different parts interact via hydrodynamic interactions. Nevertheless, using slender- body theory, Lighthill demonstrated that one can describe the hydrodynamic friction of the filament with the help of resistive force theory that introduces local friction coeffi- cients per unit length. For the translational coefficients, Lighthill obtained [31] γk = 2πη ln(2q/r) and γ⊥ = 4πη ln(2q/r) + 1/2 . (44) Reinhard Vogel, Holger Stark: Force-extension curves of bacterial flagella Here η is the shear viscosity, r = 0.02µm the cross-sectional radius of the bacterial flagellum, and q a characteristic length, for which Lighthill derived q = 0.09Λ, where Λ = 2π/√κ2 + τ 2 is the filament length of one helical turn. This gives γk ≈ 1.54η, γ⊥ ≈ 2.74η for the normal state and γk ≈ 1.64η, γ⊥ ≈ 2.91η for the coiled state. Since the coefficients are similar in both states, we use the interme- diate values γk ≈ 1.6η and γ⊥ ≈ 2.8η for simulating the force-extension curves. Finally, for the rotational friction coefficient one finds [42] γR = 4πµa2. (45) D Velocity dependence of the mean extension hζi (vp) In the following we use the numerically determined MFPT of Eq. (26) in the form hτi = τ0 exp[−β(ζ − ζ0)]. (46) We transform the probability distribution p(t) in Eq. (27) into a probability distribution for the rescaled extension variable x = β(ζ − ζ0) = βvpt/L using p(t)dt = p(x)dx. The integral in Eq. (27) is readily calculated with the help of hτi = τ0 exp(−x) and we obtain p(x)dx = ex−A exp(cid:0)−e−A(ex − 1)(cid:1) dx, (47) where exp(−A) = L/(βvpτ0). At x = 0 or ζ = ζ0, the choice of our parameters shows that the energy barrier between coiled and normal state is much larger than kBT , so p(x = 0) = e−A is almost zero or A ≫ 1. Even for x ≈ 0, we can therefore approximate the probability density in Eq. (47) as p(x) ≈ ex−A exp(cid:0)−ex−A(cid:1) dx := p0(x − A). (48) It explains why all the probability densities p(ζ, vP ) in Fig. 9 have the same shape but are shifted relative to each other along the ζ axis due to different values of vp and Γ and, therefore, of A. The distribution p(x) has a maximum at x = A with p(A) ≈ e−1. We now calculate the rescaled mean extension hxi = Z ∞ 0 xp(x)dx. (49) We rewrite the probability distribution of Eq. (48) as p(x) = ex−A exp(cid:0)−ex−A(cid:1) = −∂x exp(cid:0)−ex−A(cid:1) and obtain after integrating by parts hxi = exp(e−A)Z ∞ 0 exp(cid:0)−ex−A(cid:1) dx. (50) The substitution y = − exp(x − A) then introduces the exponential integral function: hxi = − exp(e−A)Z −e−A −∞ ey y dy, (51) Reinhard Vogel, Holger Stark: Force-extension curves of bacterial flagella 13 29. L. Landau and E. Lifshitz, Theory of Elasticity (Pergamon Press, 1986). (52) 30. S. Childress, Mechanics of swimming and flying. (Cam- bridge University Press, 1981). 31. J. Lighthill, SIAM Rev. 18, 161 (1976). 32. E.M. Purcell, Am. J. Phys 45, 3 (1977). 33. S. Chattopadhyay, R. Moldovan, C. Yeung, and X.L. Wu, PNAS 103, 13712 (2006). 34. S. Chattopadhyay and X.L. Wu, Biophys. J. 96, 2023 (2009). Reichert, Ph.D. thesis, 35. M. sity resolving.de/urn:nbn:de:bsz:352-opus-19302. Konstanz (2006), Univer- http://nbn- 36. G. Chirico and J. Langowski, Biopolymers 34, 415 (1994). 37. H. Wada and R.R. Netz, Europhys. Lett. 77, 68001 (2007). 38. H. Wada and R.R. Netz, Phys. Rev. Lett. 99, 108102 (2007). 39. D.A. Kessler and Y. Rabin, Phys. Rev. Lett. 90, 024301 (2003). 40. A. Goriely and M. Tabor, Proc. R. Soc. Lond. A, 453, 2583 (1997) 41. T.C. Flynn and J. Ma, Biophys. J. 86, 3204 (2004). 42. C. Brennen and H. Winet, Annu. Rev. Fluid Mech. 9, 339 (1977). 43. E. Evans, Annu. Rev. Biophys. Biomol. Struct. 30, 105 (2001). 44. C.W. Wolgemuth, T.R. Powers, and R.E. Goldstein, Phys. Rev. Lett., 84, 1623 (2000). 45. G. I. Bell, Science, 200, 618-627 (1978). 46. C. Dombrowski, W. Kan, M. A. Motaleb, N. W. Charon, R. E. Goldstein, and C. W. Wolgemuth, Biophys. J., 96, 4409 (2009) which for e−A ≪ 1 is approximated by hxi ≈ −(C + log e−A) = −C + A, where C ≈ 0.577 is Euler's constant. Introducing the orig- inal extension variable ζ and A = log(vpτ0/L) + log(β), we obtain for the mean extension at which the coiled-to- normal transition first takes place: β(hζi − ζ0) = −C + log(vpτ0/L) + log(β) ∝ log(vp). (53) References 1. H. C. Berg, E. coli in Motion, Springer Verlag, New York (2004). 2. L. Zhang, J.J. Abbott, L. Dong, B.E. Kratochvil, D. Bell, and B.J. Nelson, Appl. Phys. Lett. 94, 064107 (2009). 3. B. Behkam and M. Sitti, Appl. Phys. Lett. 90, 023902 (2007). 4. M.J. Kim and K.S. Breuer, Small 4, 111 (2008). 5. E. Barry, Z. Hensel, Z. Dogic, M. Shribak, and R. Olden- bourg, Phys. Rev. Lett. 96, 018305 (2006). 6. R. Kamiya and S. Asakura, J. Mol. Biol. 106, 167 (1976). 7. R. Kamiya and S. Asakura, J. Mol. Biol. 108, 513 (1977). 8. E. Hasegawa, R. Kamiya and S. Asakura, J. Mol. Biol. 160, 609 (1982). 9. H. Hotani, Biosystems 12, 325 (1980). 10. M. Seville, T. Ikeda, and H. Hotani, FEBS Lett. 332, 260 (1993). 11. R. M. Macnab and M. K. Ornston, J. Mol. Biol. 112, 1 (1977). 12. H. Hotani, J. Mol. Biol. 156, 791 (1982). 13. N. C. Darnton, L. Turner, S. Rojevsky, and H. C. Berg, J. Bacteriol. 189, 1756 (2007). 14. N. C. Darnton and H. C. Berg, Biophys. J. 92, 2230 (2007). 15. K. Hasegawa, I. Yamashita, and K. Namba, Biophys J 74, 569 (1998). 16. S. Asakura, Adv. Biophys. 1, 99 (1970). 17. C. R. Calladine, Nature 255, 121 (1975). 18. I. Yamashita, K. Hasegawa, H. Suzuki, F. Vonderviszt, Y. Mimori-Kiyosue, and K. Namba, Nat. Struct. Biol. 5, 125 (1998). 19. K. Yonekura, S. Maki-Yonekura, and K. Namba, Nature 424, 643 (2003). 20. S.V. Srigiriraju and T.R. Powers, Phys. Rev. Lett. 94, 248101 (2005). 21. S.V. Srigiriraju and T.R. Powers, Phys. Rev. E . 73, 011902 (2006). 22. B. Friedrich, J. Math. Biol. 53, 162 (2006). 23. C. Speier, Ising Model for Bacterial Flagella, diploma the- sis, Technische Universitat Berlin (2010). 24. A. Arkhipov, P. L. Freddolino, K. Imada, K. Namba, and K. Schulten, Biophys. J. 91, 4589 (2006). 25. R.E. Goldstein, A. Goriely, G. Huber, and C.W. Wolge- muth, Phys. Rev. Lett. 84, 1631 (2000). 26. D. Coombs, G. Huber, J.O. Kessler, and R.E. Goldstein, Phys. Rev. Lett. 89, 118102 (2002). 27. H. Wada and R.R. Netz, Europhys. Lett. 82, 28001 (2008). 28. A.E.H. Love, A Treatise on the Mathematical Theory of Elasticity (New York Dover Publications, 1944).
1912.04049
1
1912
2019-12-05T07:32:30
Label-free biochemical quantitative phase imaging with mid-infrared photothermal effect
[ "physics.bio-ph", "physics.optics" ]
Label-free optical imaging is valuable in biology and medicine with its non-destructive property and reduced optical and chemical damages. Quantitative phase (QPI) and molecular vibrational imaging (MVI) are the two most successful label-free methods, providing morphology and biochemistry, respectively, that have pioneered numerous applications along their independent technological maturity over the past few decades. However, the distinct label-free contrasts are inherently complementary and difficult to integrate due to the use of different light-matter interactions. Here, we present a unified imaging scheme that realizes simultaneous and in-situ acquisition of MV-fingerprint contrasts of single cells in the framework of QPI utilizing the mid-infrared photothermal effect. The fully label-free and robust integration of subcellular morphology and biochemistry would have important implications, especially for studying complex and fragile biological phenomena such as drug delivery, cellular diseases and stem cell development, where long-time observation of unperturbed cells are needed under low phototoxicity.
physics.bio-ph
physics
Label-free biochemical quantitative phase imaging with mid- infrared photothermal effect Miu Tamamitsu,1 Keiichiro Toda,1 Hiroyuki Shimada,2 Takaaki Honda,3 Masaharu Takarada,3 Kohki Okabe,3 Yu Nagashima,4 Ryoichi Horisaki,5,6 and Takuro Ideguchi1,2,6,* 1Department of Physics, The University of Tokyo, Tokyo 113-0033, Japan 2Institute for Photon Science and Technology, The University of Tokyo, Tokyo 113-0033, Japan 3Department of Pharmaceutical Sciences, The University of Tokyo, Tokyo 113-0033, Japan 4Department of Neurology, The University of Tokyo, Tokyo 113-0033, Japan 5Graduate School of Information Science and Technology, Osaka University, Osaka 565-0871, Japan 6PRESTPO, Japan Science and Technology Agency, Saitama 332-0012, Japan *Corresponding author: [email protected] Label-free optical imaging is valuable in biology and medicine with its non-destructive property and reduced optical and chemical damages. Quantitative phase (QPI) and molecular vibrational imaging (MVI) are the two most successful label-free methods, providing morphology and biochemistry, respectively, that have pioneered numerous applications along their independent technological maturity over the past few decades. However, the distinct label-free contrasts are inherently complementary and difficult to integrate due to the use of different light-matter interactions. Here, we present a unified imaging scheme that realizes simultaneous and in-situ acquisition of MV-fingerprint contrasts of single cells in the framework of QPI utilizing the mid-infrared photothermal effect. The fully label-free and robust integration of subcellular morphology and biochemistry would have important implications, especially for studying complex and fragile biological phenomena such as drug delivery, cellular diseases and stem cell development, where long-time observation of unperturbed cells are needed under low phototoxicity. Optical imaging is indispensable in biological and medical applications with its non-destructive property. Label- free imaging such as quantitative phase (QPI) and molecular vibrational imaging (MVI) is particularly valuable for studying fragile systems where exogenous labelling that often spoils the sample is not preferred1-18. QPI yields the sample-specific 2D optical-phase-delay1 or 3D refractive-index (RI) distribution2, which is the fundamental quantity used to visualize morphology of transparent samples as in the cases of the dark-field, phase-contrast and differential-interference-contrast microscopy. Compared to these traditional methods where the optical wavefront is perturbed to couple its phase into the amplitude, the direct quantification of the phase- delay or RI by QPI reveals the natural morphology that is more smooth and higher in spatial resolution and contrast, allowing for accurate and automatic cellular profiling3. Its essential capability is to translate the measured QP into the cellular dry-mass density, which has enabled quantification of cellular growth rate4 and numerous other applications5. On the other hand, MVI yields the comprehensive spectroscopic information of biomolecular bonds based on Raman scattering6-10 or mid-infrared (MIR) absorption11-18. Coherent Raman7-10 and, more recently, MIR-photothermal12-17 and -photoacoustic18 imaging have gained attention due to the high spatial resolution and detection sensitivity. The state-of-the-art systems can perform video-rate imaging of, e.g., intracellular proteins, lipids or nucleic acids, allowing for high dimensional metabolic analysis10, etc. In spite of their independent technological maturity, the inherent limitations of MVI and QPI are still left unresolved. The RI does not offer chemical specificity5 whereas the MV can be detected at limited regions where resonant molecules exist. It is also difficult to estimate the quantity of each biochemical constituent based on MV spectroscopy without the additional knowledge of the interaction lengths, the attenuation or scattering cross- sections6, and the molecular masses of unknown biological compositions. Combining the two complementary information, i.e., quantitative morphology and qualitative biochemistry, can not only mitigate these limitations, but also synergistically expand the capability of label-free imaging. Indeed, multi-modal spontaneous Raman- QPI has shown potential to decompose the local cellular dry-mass density into the individual biochemical constituents19. However, merely combining independent modalities is not a robust solution because there are always mismatches in resolutions, sampling points, fields of view (FOVs), etc. For instance, in the spontaneous Raman-QPI system, the lack of depth-resolution in the QPI prohibits accurate decomposition of the dry-mass density into independent biomolecular components, whereas the slow acquisition speed of the spontaneous Raman imaging is accompanied by motion-blur artifacts of cellular dynamics. Accurate correlation of the spatiotemporal evolutions of subcellular morphology and biochemistry yields more comprehensive and robust pictures of complex biological systems, and a fully label-free implementation would have important implications when studying fragile phenomena. Here, we present a unified imaging scheme that bridges this technological gap between the two label-free modalities, realizing simultaneous and in-situ acquisition of MV contrasts in the framework of QPI using the MIR photothermal effect. Preliminary results on this method, which we term MV-sensitive QPI (MV-QPI), have been recently reported20. The focus of this work is to further develop the MV-QPI method, proving its practical bioimaging capability in the broadband MIR fingerprint region while also pioneering the depth- and super- resolved imaging performance beyond the diffraction limit posed in other MVI techniques6-18. To highlight MV- QPI's versatility, we present two implementations of QPI scheme for the single-cell imaging application. The first is based on digital holography (DH) where we demonstrate live-cell 2D MV-QPI. The second is based on optical diffraction tomography (ODT) where we demonstrate, for the first time to our knowledge, the depth- resolved MV-QPI. We expect our MV-QPI method allows for merging of the independent knowledge of the QPI and MVI communities, leading to more comprehensive understanding of various biological phenomena. Results MV-QPI. The concept of MV-QPI is illustrated in Fig. 1. MV-QPI is a super-resolved MIR imaging method based on the wide-field visible (VIS) QPI detection of site-specific RI changes induced by the MIR photothermal effect. The entire volume of the sample placed in the objective focus of a QPI system is illuminated by MIR light of a certain wavenumber that excites the resonant molecular species to their fundamental vibrational states (see Fig. 1a). Through the non-radiative decay of the MVs, the local RIs in the vicinities of the resonant molecules decrease due to the rise of the temperature12-17, which are detected by QPI with the spatial resolution of the VIS probe light20. The obtained MIR "OFF" QPI reveals the quantitative morphology of the sample, while the subtraction of the "OFF" from the "ON" QPI reveals the site-specific phase or RI decrease which reflect the local MIR absorption property (see Fig. 1b). Scanning the MIR wavenumber yields spectroscopic images of different MV resonances. The broadband MIR absorption spectrum can be obtained at each spatial point, which can be used to identify the local molecular compositions through chemometric analysis. Eventually, we can map the MV-based biomolecular distributions within the global morphology of the sample provided by the QP contrast. Generally, the MIR photothermal imaging12-17 including our MV-QPI method20 offers (1) the high MV detection sensitivity based on the MIR absorption having ~8 orders of magnitude larger cross-section compared to Raman scattering17 and (2) the low photodamage by the use of the low-photon-energy MIR excitation that most unlikely excites the electronic transitions21 of biomolecules. In this work, we additionally harness the capability of QPI to computationally synthesize the 3D spatial-frequency aperture of the imaging system2,22. This enables us to further achieve (3) the higher lateral spatial resolution by the expanded bandwidth of the synthetic aperture that surpasses the numerical aperture (NA) of a single objective lens which poses the diffraction limit in other far- field MVI techniques6-18,20 and (4) decoupling the undesired MIR absorption effect of the surrounding aqueous medium, which is the well-known problem of MIR microscopy in general, by the depth-resolving power offered by the axial bandwidth of the 3D synthetic aperture. Figures 1c and 1d describe the mechanism of the diffraction limit when the sample is illuminated with a coherent wavefront2,22,23. Figure 1c illustrates the case of standard 2D QPI where the synthetic aperture technique is not used20. The object is illuminated with an orthogonal plane wave where a higher-frequency structure diffracts the wavefront with a larger angle. The objective lens collects the transmitted field with a limited angular range determined by its NA, posing the diffraction limit due to the low-pass filtering effect. Other MVI techniques that use focused optical probe6-18 share the same diffraction-limit mechanism posed by the NA of a single objective lens. In the synthetic-aperture QPI, this limitation can be surpassed using tilted plane wave illumination2,22,23 (see Fig. 1d). The propagation wavevector of the diffracted wavefront is accordingly tilted, such that the higher- frequency contents can be collected with the same objective lens. Essentially, the tilt shifts the location of the objective lens' NA in the spatial-frequency domain. The Fourier diffraction theorem24 also allows us to map the 2D frequency aperture to the 3D frequency space when the illumination is monochromatic, which becomes a spherical cap called the Ewald's sphere as shown in Figs. 1c and 1d. The 3D position of the spherical cap can be stirred by scanning the tilt and azimuthal angles of the illumination24, allowing us to computationally fill a certain volume of the 3D frequency space2,22. The expanded axial and lateral bandwidths of the 3D synthetic aperture result in the depth- and super-resolved imaging performance, respectively. In an extreme case, the half-pitch lateral and axial resolutions down to 90 and 150 nm have been realized, respectively22. Experimental systems. Our experimental implementations are schematically shown in Fig. 2. The VIS (10 ns duration, 532 nm wavelength) and MIR (1 µs duration, tunable wavenumber between 1,450 -- 1645 cm-1) lasers produce electrically synchronized optical pulse trains of 1 kHz repetition rate with a fixed time-delay of the former to the latter pulse (see Fig. 2a). The MIR light is intensity-modulated by a square wave such that the image sensor alternately captures the MIR ON and OFF frames at ~100 Hz. The MIR pulse fluence is ~10 pJ/µm2 (~100 nJ over ~100 µm ×100 µm) but depends on the wavenumber. The VIS pulse fluence can be as low as ~0.1 pJ/µm2 (~1 nJ over ~100 µm × 100 µm), which is 3 -- 4 orders of magnitude lower than that used in, e.g., coherent Raman imaging10. Figure 2b shows our MV-DH system. DH is a wide-field interferometric technique to measure the 2D optical- phase-delay map1,20,25. The collimated laser beam illuminates the sample and its magnified complex-field image is replicated by the subsequent diffraction grating. The zeroth-order term is low-pass filtered to create a quasi- plane reference wave while the first-order term is transmitted unperturbed, such that these two terms create an interferogram on the image sensor. The illumination optical power is ~100 µW which is enough to use the full dynamic range of the image sensor that runs at 100 Hz. The mechanism of the diffraction limit of DH corresponds to the situation illustrated in Fig. 1c. The diffraction-limited half-pitch lateral resolution is ~440 nm as determined by the objective lens NA of 0.6. The temporal phase sensitivity of our DH system is dominated by the optical shot noise and ~10 mrad in standard deviation without averaging. Figure 2c shows our MV-ODT system. ODT estimates the depth-resolved 3D RI map of the sample through the multi-angle tomographic measurements and computational reconstruction incorporating the diffraction effect2,22. The collimated VIS laser beam illuminates the sample with various incident angles stirred by the rotating wedge prism, and its magnified complex-field image is detected with the image sensor by means of Mach-Zehnder interferometry with the reference wave. We use the fixed illumination NA of 0.55 and scan 9 azimuthal angles with an increment of 36 degrees. The illumination optical power is ~1 µW which is enough to use the full dynamic range of the image sensor that runs at 60 Hz. The mechanism of the diffraction limit of ODT corresponds to the situation illustrated in Fig. 1d. The diffraction-limited half-pitch lateral and axial resolutions are ~190 nm and ~2.3 µm, respectively, as determined by the illumination and collection NAs of 0.55 and 0.85, respectively. The temporal RI sensitivity of our ODT system is dominated by the optical shot noise and ~2 × 10- 5 in standard deviation without averaging. Basic performance of MV-QPI. We first characterize the basic performance of our MV-QPI systems. The performance of our MV-DH system is summarized in our prior work20. The performance of our MV-ODT system is summarized in Fig. 3. We measure liquid oil sandwiched between two CaF2 substrates of 500 µm thickness which is excited by the MIR beam with the focus diameter of ~30 µm. We average 500 MIR ON-OFF measurements for each photothermal tomogram. In Fig. 3a, we verify that the signal (i.e., the RI decrease) varies linearly against the MIR excitation energy. In Fig. 3b, we confirm exponential temporal decay of the signal with the decay constant of ~130 µs by varying the time-delay between the MIR and VIS pulses. Similar decay constants could be obtained with other experimental conditions as the thermal diffusivities of various liquids and polymers are in the order of 10-7 [m2/s]26,27. In Fig. 3c, we perform the MIR spectroscopy of the oil by scanning the MIR wavenumber. The obtained spectrum shows good agreement with the spectrum of the same oil obtained by a commercial Fourier-transform infrared spectrometer (FTIR). Comparison of the depth-resolving effects between MV-DH and MV-ODT. In the context of MV-QPI, the depth-resolved quantitative RI imaging capability of ODT is critical for (1) decoupling the MIR photothermal effects in the out-of-focus aqueous layers and (2) quantifying the actual photothermal temperature change. We demonstrate these effects in Fig. 4 by comparing the photothermal images of fixed HEK293 cells immersed in D2O-based phosphate-buffered saline (PBS) obtained with the DH and ODT reconstruction algorithms. To maintain the consistency in the FOV and spatial resolution, we apply the synthetic-aperture DH algorithm23 on the same raw measurement dataset as those used in the ODT reconstruction. In short, the synthetic-aperture DH enhances the lateral resolution of DH by the same principle as the multi-angle ODT, but without mapping to the 3D frequency space. We average 2,500 MIR ON-OFF measurements, resulting in the acquisition time of 12.5 minutes. In this experiment, the MIR fluence is intentionally made non-uniform within the FOV to make a clearer comparison. In Fig. 4c, the cell's photothermal signals are contaminated by the non-uniform background originating from the water's photothermal phase change reflecting the MIR fluence. This background can be made more flattened by sectioning only the in-focus layer with the ODT algorithm, as shown by the blue regions of the cross-sectional profiles (see Figs. 4c and 4d). Also, some of the intracellular structures, as those indicated by the red regions in the cross-sectional profiles, are resolved with higher contrasts in the ODT reconstruction. Furthermore, ODT quantifies the photothermal RI change to be ~10-5, from which the intracellular temperature rise can be estimated to be ~0.1 K assuming the water's thermo-optic coefficient of ~1.4 × 10-4 [1/K]28. Note that the estimation of the temperature rise is generally not possible with MV-DH because the thickness and RI information are coupled in the obtained phase values. Live-cell, broadband MIR-fingerprint MV-QPI. We demonstrate live-cell, broadband MIR-fingerprint imaging of a COS7 cell immersed in H2O-based culture medium with the MV-DH system (see Fig. 5). We average 2,500 MIR ON-OFF frames for each spectral point, resulting in the acquisition time of 50 seconds. The water's MIR absorption effect is independently measured and subtracted. The QPI provided in Fig. 5a reveals the comprehensive morphology of the cell where the global cellular shape and various intracellular structures can be recognized such as nucleus, nucleoli and small cytoplasmic particles. We acquire MV-spectral images at 27 spectral points between 1,453 and 1,632 cm-1. This spectral range resides in the MIR fingerprint region where CH2 bending (1,450 -- 1,500) and peptide bond's amide II (1,500 -- 1,580) and amide I (1,580 -- 1,700) bands show characteristic spectral signatures, which are recognized to be abundant in lipids and proteins, respectively11. Figure 5b shows the obtained MIR spectrum at different locations in the FOV. We can observe different spectral signatures across different intracellular structures; e.g., compared to the cytoplasm, the nucleolus shows a stronger signal of the amide II band centered at ~1,550 cm-1. Indeed, the MV image of 1,548 cm-1 clearly visualizes the signal localizations at the nucleoli which could represent the richness of proteins (see Fig. 5d). Also, the small cytoplasmic localizations of 1,472 cm-1 signal at the cellular boundary could represent the existence of lipid droplets (see Fig. 5c). Depth-resolved, broadband MIR-fingerprint MV-QPI. We demonstrate depth-resolved, broadband MIR- fingerprint imaging of fixed HEK293 cells immersed in D2O-based PBS with the MV-ODT system (see Fig. 6). We average 1,500 MIR ON-OFF measurements for each spectral point, resulting in the acquisition time of 7.5 minutes. We acquire MV-spectral tomograms at 19 spectral points between 1,502 and 1,632 cm-1. The cross- sectional images of the sample's reconstructed 3D RI distribution at two different heights are shown in Figs. 6a and 6b, sectioning the podia (red arrow) and the nucleoli (red square) of the cells, respectively. The MIR spectrum obtained at the nucleolus resolves the spectroscopic signatures of the amide II band (see Fig. 6e) and its resonance at 1,563 cm-1 is shown in Figs. 6c and 6d. The MV contrast is localized at the nucleoli in the top two cells, while it shows a characteristic cytoplasmic distribution surrounding the nucleus in the bottom cell which reminds us of an endoplasmic reticulum. Such intercellular variation, mainly of protein distribution, could represent different phases of the cell cycle. Finally, we can observe the depth-resolved localization of the MV signal and the RI contrast originating from the nucleolus in Fig. 6f. To the best of our knowledge, this is the first demonstration of the depth-resolved MV-QPI which is enabled by the implementation of the ODT scheme. The temperature rise inside the cells can be estimated to be ~0.1 K assuming the water's thermo-optic coefficient of ~1.4 × 10-4 [1/K]. Discussions The current MV-QPI systems have rooms for improvement. First, the spatial resolution of the photothermal images is limited by the heat diffusion that happens within the relatively long MIR pulse duration29 (i.e., ~700 nm of the diffusion length in 1 µs). We estimate ~10 ns pulse duration confines the diffusion within the diffraction limit of the QPI systems. Second, a higher-energy MIR pulse laser is desired to increase the photothermal signal by an order of magnitude by creating the temperature rise of ~1 K. It is also desired to broaden the spectral tunability of the MIR light source so that various other biomolecules can be probed. In our experimental systems, these issues arise due to the use of the semiconductor MIR quantum cascade laser that offers low optical output power with a limited gain bandwidth. Third, a higher-full-well-capacity image sensor30 can be used to enhance the shot-noise-limited sensitivities of our MV-QPI systems by at most two orders of magnitude. Fourth, any other QPI methods can be implemented to harness their versatility5, robustness31-33 and imaging speed34-37 while higher-NA objective lenses can be used to enhance the lateral and axial spatial resolutions22. With the above-mentioned improvements, MV-QPI method has the potential to achieve high-sensitive, low- photodamage and super-resolution MVI. In total, nearly three orders of magnitude enhancement in the MV detection sensitivity can be expected. In this case, the VIS fluence at the sample plane increases by ~3 orders of magnitude and becomes comparable to that used in coherent Raman imaging10. Our method can still reduce photodamage associated with nonlinear electronic transitions of biomolecules21 since the duration of the VIS pulse can be ~10 ns which is significantly longer than the picosecond or femtosecond pulses used in coherent Raman imaging. Also, with the reduced MIR pulse duration, the spatial resolution of the photothermal images can, in principle, reach diffraction limit of the synthetic-aperture QPI system, which surpasses the diffraction limit posed in other MVI techniques6-18,20 by the NA of a single objective lens. References 1. Marquet, P. et al. Digital holographic microscopy: a noninvasive contrast imaging technique allowing quantitative visualization of living cells with subwavelength axial accuracy. Opt. Lett. 30, 468-470 (2005). 2. Sung, Y. et al. Optical diffraction tomography for high resolution live cell imaging. Opt. Express 17, 266-277 (2009). 3. Kasprowicz, R., Suman, R. & O'Toole, P. Characterizing live cell behavior: Traditional label-free and quantitative phase imaging approaches. Int. J. Biochem. Cell Biol. 84, 89-95 (2017). 4. Popescu, G. et al. Optical imaging of cell mass and growth dynamics. Am. J. Physiol. Cell Physiol. 295, C538-C544 (2008). 5. Park, Y., Depeursinge, C. & Popescu, G. Quantitative phase imaging in biomedicine. Nat. Photonics 12, 578-589 (2018). 6. Shipp, D. W., Sinjab, F. & Notingher, I. Raman spectroscopy: techniques and applications in the life sciences. Adv. Opt. 7. Camp Jr., C. H. & Cicerone, M. T. Chemically sensitive bioimaging with coherent Raman scattering. Nat. Photonics 9, 295-305 8. Cheng, J. X. & Xie, X. S. Vibrational spectroscopic imaging of living systems: An emerging platform for biology and medicine. 9. Hu, F., Shi, L. & Min, W. Biological imaging of chemical bonds by stimulated Raman scattering microscopy. Nat. Methods 16, 10. Wakisaka, Y. et al. Probing the metabolic heterogeneity of live Euglena gracilis with stimulated Raman scattering microscopy. Photonics 9, 315-428 (2017). (2015). Science 350, aaa8870 (2015). 830-842 (2019). Nat. Microbiol. 1, 16124 (2016). 11. Baker, M. J. et al. Using Fourier transform IR spectroscopy to analyze biological materials. Nat. Protoc. 9, 1771-1791 (2014). 12. Lim, J. M. et al. Cytoplasmic Protein Imaging with Mid-Infrared Photothermal Microscopy: Cellular Dynamics of Live Neurons and Oligodendrocytes. J. Phys. Chem. Lett. 10, 2857-2861 (2019). 13. Bai, Y. et al. Ultrafast chemical imaging by widefield photothermal sensing of infrared absorption. Sci. Adv. 5, eaav7127 (2019). 14. Toda, K., Tamamitsu, M., Nagashima, Y., Horisaki, R. & Ideguchi, T. Molecular contrast on phase-contrast microscope. Sci. Rep. 9, 9957 (2019). 15. Samolis, P. D. & Sander, M. Y. Phase-sensitive lock-in detection for high-contrast mid-infrared photothermal imaging with sub- diffraction limited resolution. Opt. Express 27, 2643-2655 (2019). 16. Li, Z., Aleshire, K., Kuno, M. & Hartland, G. V. Super-Resolution Far-Field Infrared Imaging by Photothermal Heterodyne 17. Zhang, D. et al. Depth-resolved mid-infrared photothermal imaging of living cells and organisms with submicrometer spatial Imaging. J. Phys. Chem. B 121, 8838-8846 (2017). resolution. Sci. Adv. 2, e1600521 (2016). 18. Shi, J. et al. High-resolution, high-contrast mid-infrared imaging of fresh biological samples with ultraviolet-localized photoacoustic microscopy. Nat. Photonics 13, 609-615 (2019). 19. Pavillon, N., Hobro, A. J. & Smith, N. I. Cell Optical Density and Molecular Composition Revealed by Simultaneous Multimodal Label-Free Imaging. Biophys. J. 105, 1123-1132 (2013). 20. Tamamitsu, M., Toda, K., Horisaki, R. & Ideguchi, T. Quantitative phase imaging with molecular vibrational sensitivity. Opt. 25. Bhaduri, B., Pham, H., Mir, M. & Popescu, G. Diffraction phase microscopy with white light. Opt. Lett. 37, 1094-1096 (2012). 26. Wang, J. & Fiebig, M. Measurement of the thermal diffusivity of aqueous solutions of alcohols by a laser-induced thermal grating technique. Int. J. Thermophys. 16, 1353-1361 (1995). 27. Salazar, A. On thermal diffusivity. Eur. J. Phys. 24, 351-358 (2003). 28. Schiebener, P., Straub, J., Levelt Sengers, J. M. H. & Gallagher, J. S. Refractive index of water and steam as function of wavelength, temperature and density. J. Phys. Chem. Ref. Data 19, 677 -- 717 (1990). 29. Zharov, V. P. & Lapotko, D. O. Photothermal Imaging of Nanoparticles and Cells. IEEE J. Sel. Top. Quant. 11, 733-751 (2005). 30. Hosseini, P. et al. Pushing phase and amplitude sensitivity limits in interferometric microscopy. Opt. Lett. 41, 1656-1659 (2016). 31. Chowdhury, S. et al. High-resolution 3D refractive index microscopy of multiple-scattering samples from intensity images. 32. Merola, F. et al. Tomographic flow cytometry by digital holography. Light Sci. Appl. 6, e16241 (2017). 33. Kim, K. & Park, Y. Tomographic active optical trapping of arbitrarily shaped objects by exploiting 3D refractive index maps. Optica 6, 1211-1219 (2019). Nat. Commun. 8, 15340 (2017). 34. Horisaki, R., Fujii, K. &Tanida, J. Diffusion-based single-shot diffraction tomography. Opt. Lett. 44, 1964-1967 (2019). 21. Fu, Y., Wang, H., Shi, R. & Cheng, J. X. Characterization of photodamage in coherent anti-Stokes Raman scattering microscopy. 22. Cotte, Y. et al. Marker-free phase nanoscopy. Nat. Photonics 7, 113-117 (2013). 23. Micó, V., Zheng, J., Garcia, J., Zalevsky, Z. & Gao, P. Resolution enhancement in quantitative phase microscopy. Adv. Opt. 24. Wolf, E. Three-dimensional structure determination of semi-transparent objects from holographic data. Opt. Commun. 1, 153-156 Lett. 44, 3729-3732 (2019). Opt. Express 14, 3942-3951 (2006). Photonics 11, 135-214 (2019). (1969). 36. Rodrigo, J. A., Soto, J. M. & Alieva. T. Fast label-free microscopy technique for 3D dynamic quantitative imaging of living cells. tomography. Opt. Lett. 42, 999-1002 (2017). Biomed. Opt. Express 8, 5507-5517 (2017). 35. Lee., K., Kim, K., Kim, G., Shin, S. & Park, Y. Time-multiplexed structured illumination using a DMD for optical diffraction 37. Horisaki, R., Egami, R. & Tanida, J. Single-shot phase imaging with randomized light (SPIRaL). Opt. Express 24, 3765-3773 (2016). Methods Light sources. The VIS light source is the second harmonic generation (SHG) of 10 ns, 1 kHz, 1,064 nm pulsed laser beam emitted from the Q-switch laser NL204-1K (Ekspla) produced with the nonlinear crystal LBO-503 (Eksma Optics). The spatial mode of the SHG beam is cleaned by the single-mode optical fiber P3-405B-FC-5 (Thorlabs). The MIR light source is DO418 (Hedgehog, Daylight Solutions) providing access to 1,450 - 1,640 cm-1 region with the pulse duration of ~1 µs. Measurement of the MIR pulse-energy spectrum. The intensity spectrum of the MIR light source is generally not uniform; hence, the raw photothermal (i.e., MIR ON-OFF) contrast needs to be normalized by the corresponding MIR pulse energy in order to obtain the sample-specific spectroscopic information. We use a cadmium-mercury telluride detector PVM-10.6-1x1-TO8 (VIGO System) to measure the waveform of the MIR pulse at each wavenumber with an oscilloscope TDS2024C (Tektronix), and use the area of the obtained waveform as a measure of the pulse energy. This measurement is performed independent of the image acquisition. MV-DH system. The image sensor is operated at 100 Hz with the exposure time of 9 ms. The MIR ON-OFF modulation rate is 50 Hz. The VIS illumination power at the sample plane is ~100 µW which is enough to use the full dynamic range of the image sensor. The MIR pulse energy at the sample plane is ~100 nJ on average but depends on the wavenumber (e.g., ~200 nJ at 1,548 cm-1). MV-ODT system. The image sensor is operated at 60 Hz with the exposure time of 15 ms. The MIR ON-OFF modulation rate is 30 Hz. The VIS illumination power at the sample plane is ~1 µW which is enough to use the full dynamic range of the image sensor. The MIR pulse energy at the sample plane is ~100 nJ on average but depends on the wavenumber (e.g., ~200 nJ at 1,548 cm-1). Materials used to characterize the basic performance of the MV-ODT system. Series A 1.54000 (Cargille) is used as the liquid oil sample. FT/IR-6800, ATR PRO ONE and PKS-D1F (JASCO) are used to obtain the reference MIR absorption spectrum of the oil. Biological samples. The COS7 cells (Riken) are cultured in Dulbecco's Modified Eagle's Medium (DMEM) with 10% fetal bovine serum supplemented with penicillin -- streptomycin, L-glutamine, sodium pyruvate and nonessential amino acids at 37 °C in 5% CO2. For the live-cell imaging, the cells are cultured in 35-mm glass- bottomed dishes (AGC Techno Glass) and the medium is replaced by phenol red-free culture medium containing HEPES buffer (2 mL) before imaging. All solutions are from Thermo Fisher Scientific. The HEK293 cells are fixed with 4% paraformaldehyde at room temperature for 5 minutes and immersed in D2O-based PBS before observation. All the samples are sandwiched between two CaF2 substrates of 500 µm thickness before observation. Data availability The data provided in the manuscript are available from T.I. upon request. Acknowledgements We thank Makoto Kuwata-Gonokami and Junji Yumoto for letting us use their equipment. This work was financially supported by JST PRESTO (JPMJPR17G2) and JSPS KAKENHI (17H04852, 17K19071). Author contributions T.I. conceived the concept. M.T. proposed MV-ODT, designed and constructed the systems, designed and performed the experiments and analyzed the experimental data. K.T. contributed in design of the systems and the experiments, interpretation of the obtained data and daily discussion regarding the MV-QPI method. H.S. wrote the automated data acquisition software program and the graphical user interface. T.H., M.K. and K.O. prepared the COS7 cell. Y.N. prepared the HEK293 cells and provided biological perspectives for the data interpretation. R.H. advised on the computational reconstruction framework of ODT. T.I. supervised the entire work. M.T. and T.I. wrote the manuscript with contributions from all the other authors. Competing interests Authors declare no competing interest. Fig. 1 Concept of MV-QPI. a, Principle of the MV-contrast acquisition in the QPI framework. The MIR light of a certain wavenumber is irradiated to the entire volume of the sample, where the resonant biomolecules are selectively excited to their fundamental vibrational states. The vibrational energy is eventually transformed into heat that diffuses into the surrounding medium. The resulting photothermal RI decrease is detected by the QPI system with the spatial resolution of the VIS probe light. b, Cross-correlative analysis enabled by MV-QPI. The phase or RI image obtained at the MIR OFF state reveals the quantitative and comprehensive morphology of the sample containing rich information about cellular shapes and intracellular-organelle distributions. Scanning of the MIR wavenumber visualizes contrasts of various MV resonances at each spatial point of the FOV, which can be decomposed into individual biomolecular constituents through chemometric analysis. c, Mechanism of the diffraction limit in the standard 2D QPI. The object is illuminated with the orthogonal plane wave and only a limited range of spatial frequency information of the diffracted field can be collected with the NA of the objective lens. d, Mechanism of the depth- and super-resolution in the synthetic-aperture QPI. The object is illuminated with the tilted plane wave such that higher-frequency contents can be collected that used to be outside the NA of the objective lens in c. Scanning the tilt and azimuthal angles of the illumination allows us to computationally synthesize the 3D frequency aperture. The depth- and super-resolved imaging performance can be achieved with the expanded lateral and axial bandwidths of the 3D synthetic aperture, respectively. The black dotted curves in the frequency spectrum indicate the Ewald's spherical cap that determines the 3D coverage of the objective lens' NA under a certain angle of illumination. MIR excitationsampleQPI systemphaseor RIMIR OFFphase or RIdecreaseMIR ν1: ON-OFF phase or RIdecreaseMIR ν2: ON-OFFνMIR wavenumbersignalnucleicacidlipidproteinlipidabvibrationalresonanceRI changeby heatbiomoleculeCNOHRR12orthogonal illuminationtilted illumination expandedbandwidth(superresolution)axialbandwidth(depthresolution)kxkykxkzhigherfreqeuncycontentscdobjective lensobjective lensfrequencyspectrum Fig. 2 Experimental implementations. a, Temporal synchronization. The VIS and MIR lasers are electrically controlled to synchronize their pulse repetitions (~1 kHz) and relative time-delay. The MIR beam is intensity- modulated to be in-phase with the half-harmonic of the image sensor's frame rate (~100 Hz). b, MV-DH system. The DH microscope is built based on a commercial microscope housing IX73 (Olympus). The collimated VIS laser beam is used as the plane-wave probe illumination and the magnified image of the sample is formed next to the output port of IX73. The subsequent 4f system is used to perform the common-path off-axis interferometry. The MIR laser beam is loosely focused onto the sample with a CaF2 lens. QCL: quantum cascade laser. c, MV- ODT system. The collimated VIS laser beam is split into two paths to create the Mach-Zehnder off-axis interferometer. The deflection angle of the probe beam created by the wedge prism is magnified and relayed into the sample plane by the subsequent tube lens and the illumination objective lens. The resulting tilted plane-wave illumination is then collected by another objective lens and the subsequent tube lens forms the sample's magnified image on the image sensor. The MIR and VIS beams are combined by the dichroic mirror (DM). gratingpinholeimagesensoratimesensorexposureMIRexposureVISexposureMIR ONframeMIR OFFframeVISlight sourcebeamblockbsamplerotatingwedgeprismdelay lineetc.imagesensorVISlight sourceMIRQCLcOlympusIX73sampleDMMIRQCLtimedelay measurement y = -0.932 x -1 -0.5 a ) 4 - 0 1 ( e g n a h c I R 0 0 measurement y = -0.782 exp(-x/0.131) c ) . u . d e z i l a m r o n a ( e g n a h c I R 6 4 2 0 b ) 4 - 0 1 ( e g n a h c I R -1 -0.8 -0.6 -0.4 -0.2 0 0.2 MV-ODT FTIR 0 20 40 60 1,640 a b s o r p t i o n ( % ) 2 4 6 8 10 0 0.2 0.4 0.6 0.8 1 time delay (ms) MIR excitation energy (a.u.) Fig. 3 Basic performance of the MV-ODT system. The liquid oil sandwiched between two CaF2 substrates is used as the sample which is excited by the MIR beam with the focus diameter of ~30 µm. a, Linearity of the photothermal RI change with respect to the MIR excitation pulse energy. b, Exponential temporal decay of the photothermal RI change with the decay constant of ~130 µs. c, MIR spectrum of the liquid oil obtained by the MV-ODT system, showing good agreement with the FTIR reference spectrum. Each measurement point shown in a -- c represents one voxel of the FOV. 1,460 1,520 1,580 wavenumber (cm )-1 MIR OFF a H D e r u t r e p a - c i t e h t n y s T D O 5 μm b 5 μm MIR ON-OFF (1,548 cm )-1 c phase change (mrad) phase (rad) 2.5 2 1.5 1 0.5 0 -0.5 -6 -5 -4 -3 -2 -1 -5 -4 -3 -2 -1 0 5 μm RI d 1.344 1.340 1.326 1.332 5 μm RI change (10 )-5 -1.6 -1.2 -0.8 -0.4 0 0.4 -2 -1.5 -1 -0.5 0 -0.5 -1 0 10 20 30 40 50 60 distance (μm) 0 10 20 30 40 50 60 distance (μm) Fig. 4 Comparison of the depth-resolving effects between MV-DH and MV-ODT. a, b, Raw phase and RI images of the fixed HEK293 cells in D2O-based PBS at the MIR OFF state, respectively. b sections one particular height of the reconstructed 3D RI tomogram. c, d, Photothermal contrasts of the same FOVs as those shown in a and b, respectively, obtained with the MIR wavenumber tuned to 1,548 cm-1. In c, the cellular structures are contaminated by the photothermal signals originating from the out-of-focus aqueous layers. In d, the depth-resolution provided by ODT results in the higher contrasts of the cellular structures (indicated by the red arrows) as well as the more uniform and flattened background distribution originating from the MIR absorption of the in-focus water layer (indicated by the blue arrows). The depth-resolved quantification of the RI values also allows for accurate estimation of the photothermal temperature rise inside the cells (~0.1 K) using the thermo-optic coefficient of water (~1.4 ×10-4 [1/K]). In this experiment, the MIR fluence is intentionally made non-uniform within the FOV to make a clearer comparison. Fig. 5 Live-cell, broadband MIR-fingerprint MV-DH microscopy. a, Raw phase image of the live COS7 cell in H2O-based culture medium at the MIR OFF state. b, MIR spectrum of the nucleolus (orange), cytoplasm (blue) and empty area (gray) indicated by the arrows of the respective colors in a. The scanned MIR wavenumber region resides in the MIR fingerprint region where spectroscopic signatures of CH2 bending and peptide bond's amide I and II bands can be found, which are abundant in lipids and proteins, respectively. Compared to the cytoplasm, the nucleolus shows the stronger signal of the broad absorption centered at ~1,550 cm-1 which coincides with the amide II band. Each spectral points represents the spatial average of 3 × 3 diffraction-limited pixels (1.3 µm × 1.3 µm). c, d, MV images of the cell resonant to 1,472 and 1,548 cm-1, respectively, after normalization. In c, the small cytoplasmic localizations of the MV contrast at the cellular boundary could represent the existence of lipid droplets. In d, the MV contrast is selectively strong on the nucleoli which could represent the richness of proteins. MIR wavenumber (cm )-1amide IInormalizedphotothermal signal (a.u.)amide Iba, MIR OFF phased, 1,548 cm21.510.508765432-1a.u.radCOS75 μm5 μm65430a.u.-15 μmc, 1,472 cm-1-1CH2bending2111,4501,5001,5501,6001,65001234560 Fig. 6 Depth-resolved, broadband MIR-fingerprint MV-ODT microscopy. a, b, Cross-sectional images in two different axial planes of the reconstructed RI tomogram of the fixed HEK293 cells in D2O-based PBS at the MIR OFF state. a sections the podia of the cell (red arrow) while b the nucleoli (red square). c, d, MV contrasts of the same FOVs as those shown in a and b, respectively, resonant to 1,563 cm-1. The photothermal temperature rise inside the cells can be estimated to be ~0.1 K using the thermo-optic coefficient of water. e, MIR spectrum at one voxel of the FOV in the nucleolus indicated by the white arrow in d, resolving the characteristic signature of the amide II band. f, Enlargement of the red-square regions in a -- d. At z = 0 µm, the nucleolus indicated by the red arrows are not visible in the RI or the photothermal contrast. At z = 3.3 µm, the nucleolus appears in the RI contrast which also gives the signal in the photothermal contrast, demonstrating the depth-resolving capability. 1,5001,5401,5801,62000.511.522.533.544.55wavenumber (cm )-15 μmnormalized signal (a.u.)ea, MIR OFF RI c, 1,563 cm b, MIR OFF RI d, 1,563 cm -11.331.3351.341.3451.350.50-0.5-1-1.5-2-2.51.331.3351.341.3451.35z = 3.3 μm z = 0 μm MIR OFFMIR OFF1,563 cm -11,563 cm -1fRIRI change (10 )-5RI-1amid IIamid I5 μm5 μm5 μmz = 0 μmz = 3.3 μmz = 0 μmz = 3.3 μm0.50-0.5-1-1.5-2-2.5RI change (10 )-5
1611.02016
1
1611
2016-11-07T12:21:05
Small-angle scattering studies of intrinsically disordered proteins and their complexes
[ "physics.bio-ph", "q-bio.BM" ]
Intrinsically Disordered Proteins (IDPs) perform a broad range of biological functions. Their relevance has motivated intense research activity seeking to characterize their sequence/structure/function relationships. However, the conformational plasticity of these molecules hampers the application of traditional structural approaches, and new tools and concepts are being developed to address the challenges they pose. Small-Angle Scattering (SAS) is a structural biology technique that probes the size and shape of disordered proteins and their complexes with other biomolecules. The low-resolution nature of SAS can be compensated with specially designed computational tools and its combined interpretation with complementary structural information. In this review, we describe recent advances in the application of SAS to disordered proteins and highly flexible complexes and discuss current challenges.
physics.bio-ph
physics
Small-Angle Scattering Studies of Intrinsically Disordered Proteins and their Complexes Tiago N. Cordeiroa, Fátima Herranz-Trilloa,c, Annika Urbaneka, Alejandro Estañaa,b, Juan Cortésb, Nathalie Sibillea, Pau Bernadóa,* a-Centre de Biochimie Structurale. INSERM U1054, CNRS UMR 5048, Université de Montpellier. 29, rue de Navacelles, 34090-Montpellier, France. b- LAAS-CNRS, Université de Toulouse, CNRS, Toulouse, France. c- Department of Pharmacy and Department of Drug Design and Pharmacology, University of Copenhagen, Universitetsparken 2, 2100 Copenhagen, Denmark. Contact Author: Pau Bernadó ([email protected]). Tel. +33 467417705 1 Abstract: Intrinsically Disordered Proteins (IDPs) perform a broad range of biological functions. Their relevance has motivated intense research activity seeking to characterize their sequence/structure/function relationships. However, the conformational plasticity of these molecules hampers the application of traditional structural approaches, and new tools and concepts are being developed to address the challenges they pose. Small-Angle Scattering (SAS) is a structural biology technique that probes the size and shape of disordered proteins and their complexes with other biomolecules. The low-resolution nature of SAS can be compensated with specially designed computational tools and its combined interpretation with complementary structural information. In this review, we describe recent advances in the application of SAS to disordered proteins and highly flexible complexes and discuss current challenges. 2 Introduction In the last two decades, Intrinsically Disordered Proteins or Regions (IDPs/IDRs) have emerged as fundamental molecules in a broad range of crucial biological functions such as cell signaling, regulation, and homeostasis [1,2,3**]. Due to their lack of a permanent secondary and tertiary structure, IDPs and IDRs are highly plastic and have the capacity to perform specialized functions that complement those of their globular (folded) counterparts [4]. Disordered regions, which can finely adapt to the structural and chemical features of their partners, are very well suited for protein-protein interactions and are thus abundant in hub positions of interactomes [5,6,7]. The importance of disordered proteins in a multitude of biological processes has fostered intense research efforts that seek to unravel the structural bases of their function. Nuclear Magnetic Resonance (NMR) has been the main structural biology technique used to characterize the conformational preferences at residue level, and, therefore, to localize partially structured elements [8,9]. However, a number of structural features related to the overall size and shape of IDPs or their complexes remain elusive to NMR. To study these properties, thereby complementing NMR residue-specific information, Small-Angle Scattering (SAS) of X-rays (SAXS) or Neutrons (SANS) is the most appropriate technique [10,11,12]. Although SAS is a low-resolution technique, the data obtained is sensitive to large-scale protein fluctuations and the presence of multiple species and/or conformations in solution [13,14,15]. However, the conversion of SAS properties into structural restraints is challenging due to the enormous conformational variability of IDPs and the ensemble- averaged nature of the experimental data [16]. The quantitative analysis of these data in terms of structure has prompted the development of computational approaches to both model disordered proteins and to use ensembles of conformations to describe the experimental data. Here we highlight the most relevant developments and applications of SAS to IDPs and IDRs, with a special emphasis on the computational strategies required to fully exploit the data in order to achieve biologically insightful information. Structural models of IDPs and their experimental validation For disordered proteins, the structural insights gained from overall SAS parameters, such as the radius of gyration, Rg, the pairwise intramolecular distance distribution, p(r), and the maximum intramolecular distance, Dmax, are limited. Neither these parameters nor the 3 traditional Kratky representation, I(s)s2 vs s, which qualitatively report on the compactness of biomolecules in solution, directly account for the ensemble nature of disordered proteins. In order to fully exploit the structural and dynamic information encoded in SAS data, it is necessary to use realistic three-dimensional (3D) models. However, the generation of conformational ensembles of disordered proteins is extremely challenging, mainly because of the flat energy landscape and the large number of local minima separated by low-energy barriers [17]. The most popular methods to generate 3D models of IDPs are based on residue-specific conformational landscapes derived from large databases of crystallographic structures [18,19,20*]. However, the main limitation of these approaches is the absence of sequence context information, thereby precluding the prediction of transiently formed secondary structure elements or the presence of long- range interactions between distant regions of the protein. Accurate energy models (force- fields) accounting for the interactions within the chain and with the solvent are required to describe these features. The development of specific force-fields to study conformational fluctuations in disordered proteins is a very active field of research [21,22,23,24]. Molecular Dynamics (MD) or Monte-Carlo (MC) simulations, when an appropriate energy description is provided, are suitable methods to correctly sample the conformational space of IDPs. However, the high-dimensionality and the breadth of the energy landscape hamper exhaustive exploration of this space. Replica Exchange MD (REMD) [25,26], which exchanges conformations between parallel simulations running at multiple temperatures, or Multiscale Enhanced Sampling (MSES) [ 27 ], which couples temperature and Hamiltonian replica exchange, have been proposed to enhance the conformational exploration of MD methods. The performance of MD-based methods can also be improved by the inclusion of experimental data to delimit the exploration to the most relevant regions of the conformational space [28,29,30]. The quality of computational models of disordered proteins is normally validated using experimental data. The Rg derived from the low-angle region of SAXS curves or from the p(r) function is an excellent probe of the overall size of a particle in solution. Rg compilations have been extensively used to validate models of denatured and natively disordered proteins through Flory's relationship, which correlates the Rg observed with the number residues of the chain [31,14]. The compilation of the Rgs from 76 IDPs (Figure 1) reveals that these proteins are more compact than chemically denatured ones. It has been shown that denatured proteins present an enhanced sampling of extended conformations, 4 probably due to the interaction of the protein with chemical agents [32]. Importantly, deviations from the expected Rg values for canonical random-coil behavior, which is represented by the green line in figure 1, indicate the presence of structural features that modify the overall size of the particle in solution towards more extended or more compact (Figure 1). The extendedness detected using this analysis for several Tau protein constructs has been linked to the presence of secondary structural elements probed by NMR [33]. These structural properties can be more thoroughly examined when the complete SAXS curve is used to validate the ensemble models of peptides [34] or proteins [19,35,36]. Ensemble approaches In the last decade, ensemble methods have become highly popular to structurally characterize disordered proteins. Guided by experimental data, these methods aim to derive accurate ensemble models of flexible proteins. Several strategies that apply these methods to SAS data have been reported: Ensemble Optimization Method (EOM) [37,38]; Minimal Ensemble Search (MES) [39]; Basis-Set Supported SAXS (BSS-SAXS) [40]; Maximum Occurrence (MAX-Occ) [41]; Ensemble Refinement of SAXS (EROS) [42]; Broad Ensemble Generator with Re-weighting (BEGR) [43]; and Bayesian Ensemble SAXS (BE-SAXS) [44]. These methods share a common strategy that consists of the following three consecutive steps: (i) computational generation of a large ensemble that describes the conformational landscape of the protein; (ii) calculation of the theoretical SAXS curves from the individual conformations; and (iii) use of a multiparametric optimization method to select a sub-ensemble of conformations that collectively describe the experimental profile. Despite the common strategy, these approaches present distinct features in the three steps. Readers are referred to the original articles for detailed descriptions. The availability of ensemble methods has transformed the study of flexible proteins by SAS. Ensemble methods provide a description in terms of the statistical distributions of structural parameters or conformations that is revolutionary with respect to traditional analyses based on averaged parameters extracted from raw data. Using this power, structural perturbations exerted by temperature [45*,46], buffer composition [47], or mutations [48] have been monitored in terms of ensembles of conformations. 5 Despite the popularity of ensemble methods, several aspects are still under debate. The most relevant ones are the use of discrete descriptions for entities that probe an astronomical number of conformations, and the statistical significance of ensembles derived from data containing a very limited amount of information. The strategies described use distinct philosophies to address these issues, including the search for the minimum number of conformations to describe the data [38,39], the representation of the optimal solution as a distribution of low-resolution structural parameters such as Rg or Dmax [37], and the application of Bayesian statistics [40,44] or maximum entropy approaches [42]. Regardless of the strategy used to derive an ensemble of conformations compatible with the experimental data, one must be careful on the structural interpretation of the final solution. The optimized ensemble is a representation of the behavior of the protein in solution and not the exact enumeration of the conformations adopted by the protein. Consequently, the final ensemble can only be used to derive structural features that describe the protein. Importantly, the nature of these features depends on the experimental data used to derive the model. If only SAS data have been used, then an assessment of the degree of flexibility, and the size and shape distributions sampled by the protein can be obtained from the ensemble. Conversely, conformational preferences at residue level can be extracted if NMR information probing structure in a residue-specific manner is used along the refinement. Enriching the definition of conformational ensembles of IDPs with complementary information The definition of protein ensembles derived from SAS data using ensemble methods is limited to the overall structure and the space sampled by the protein in solution. Although this is an important improvement with respect to classical approaches, several crucial features, such as the localization of secondary structural elements or compact regions, remain elusive using this approach. Considerable research efforts have been channeled into enriching the resolution of the resulting ensemble with complementary information. NMR is the only technique that can provide atomic-resolution information on IDPs and, consequently, it is the most common method applied in combination with SAS [49]. NMR is highly versatile and can measure multiple observables reporting on protein structure and dynamics [50]. Concretely, information reporting on the backbone conformational 6 preferences at residue level can be probed by means of time- and ensemble-averaged chemical-shifts (CSs), J-couplings and Residual-Dipolar Couplings (RDCs). NMR can also probe long-range interactions within a protein chain or in protein complexes through Paramagnetic Relaxation Enhancement (PRE) experiments. In these experiments, a stable radical or a paramagnetic metal is introduced in a specific position of the chain, and the spatially close atoms can be identified by a decrease in their signal intensity that is proportional to the distance. The best manner to exploit the complementarity between NMR and SAS is to integrate the experimental data into the same refinement protocol. The programs ENSEMBLE [51,52] and ASTEROIDS [53] derive ensembles of disordered proteins by collectively describing SAXS curves, in addition to several NMR observables. These powerful approaches seek to find the appropriate way to combine data with very different information content while avoiding overfitting. In a pioneering study, ensembles of Tau and α-synuclein were determined by combining SAXS with multiple backbone CS, RDC, and PRE datasets [54**]. Those authors addressed the optimal combination of experimental data and the overfitting problem with extensive cross-validation tests that substantiated conformational bias in the aggregation-nucleation regions for both proteins. Other structural techniques such as single molecule Fluorescence Resonance Energy Transfer (smFRET) [55] and Electron Paramagnetic Resonance (EPR) [56,57**] have been combined with SAXS to study large and flexible complexes. Recent developments in Mass Spectrometry (MS) offer novel sources of structural information [58]. Ion Mobility Spectrometry (IMS) can capture, in a similar way to SAS, the overall properties of conformational ensembles of disordered proteins. However, a recent study comparing IMS and SAXS data for some IDPs suggests that the conformations sampled in solution and in gas-phase are not equivalent [59]. Hydrogen/Deuterium Exchange MS (HDX/MS) probes structural elements in proteins by identifying regions that are protected from the exchange with solvent protons [58]. The availability of fast HDX/MS methods enables the exploration of secondary structural elements in IDPs and localizing their interaction sites with globular partners [60]. In a recent study HDX/MS information was combined with SAXS to study the calcium-induced structure formation in RD, a protein hosting repeated regions able to bind this cation [61]. 7 The structural definition of a SAXS derived ensemble model can also be enriched by the simultaneous analysis of curves measured for multiple deletion mutants of the same IDP [37]. When applied to two different isoforms of Tau protein, this approach identified the repeat region of the protein as the origin of distinct global rearrangements of its flanking regions [62]. The large toolbox of structural techniques that can probe distinct structural features of IDPs will result in a better understanding on their structure-function relationship. In this regard, the future development of robust and reliable ways to integrate biophysical measurements in ensemble approaches is imperative when addressing complex biomolecular entities such as IDPs and their complexes. Disordered proteins in complexes The biological function of many IDPs is manifested when they recognize their biological folded partners [5]. This recognition frequently involves linear motifs of the disordered chain, which, upon binding, adopt relatively fixed conformations while the rest of the IDP remains flexible [63]. The relevance of protein-protein complexes involving disordered partners has promoted growing interest in unraveling their structural characterization, with the aim to understand the bases of their biological activity. This structural characterization is complex and poses multiple challenges to traditional structural biology methods. SAXS has emerged as a valuable alternative. However, overall structural parameters or ab initio reconstructions derived from SAXS curves cannot capture the inherent plasticity of these complexes [64,65*,66]. Hybrid (or integrative) methods that combine information from multiple techniques, thus exploiting their individual strengths, are the most appropriate approaches to study highly flexible complexes [67]. In this context, it is important to describe how different structural biology techniques probe complexes involving IDPs (Figure 2). Due to the dynamic nature of the interaction and the distinct hydrodynamic properties of the globular and disordered parts of the complex, NMR generally detects only those regions that remain flexible upon binding. Although not general, it is sometimes possible to crystallize the globular partner in the presence of a small peptide corresponding to the interacting region of the IDP. Therefore, X-ray crystallography provides an atomic resolution picture of the interacting regions that is 8 complementary to NMR since the two techniques probe non-overlapping parts of the same entity [68]. Conversely, SAXS probes the complete assembly and can be used to integrate the information from both NMR and X-ray crystallography. If one of the partners is deuterated, contrast variation SANS experiments can be performed and the individual components of the assembly can be alternatively highlighted depending on the D2O/H2O ratio of the buffer. The power of combining multiple techniques is exemplified in the study of the interaction of the Vesicular Stomatitis Virus (VSV) nucleoprotein (N0) and the dimeric phosphoprotein (P), a high-affinity complex that precludes the oligomerization of N0 in vivo [ 69 **]. Using EOM, the authors simultaneously fitted one SAXS curve and four SANS curves measured at different contrast levels for the complex of N0 with deuterated P protein. The additional information provided by the distinct contribution of the two proteins in the SANS experiments notably improved the description of the conformational properties of the complex. In many cases, the conformational mobility of the interacting region of the IDP is reduced (or frozen) upon binding to the biological partner. There is an entropic cost associated with this rigidification that often leads to low- to moderate-affinity complexes (Kd > 1 μM) [63]. The structural modulation of the affinity is key to achieving tunable responses to external signals, thereby explaining the prevalent role of disordered proteins in signaling processes [2,3**]. In the concentration range normally used in SAXS experiments, the complex is in equilibrium with the free forms of the two partners, thereby giving rise to population-weighted averaged SAXS curves (Figure 3A). This scenario can be even more complex if one or both of the partners have multiple equivalent or similar binding sites (Figure 3B,C). In this case, the polydispersity of the mixture increases as a result of the presence of several complexes with distinct stoichiometries. The interpretation of SAS data from polydisperse samples is challenging [70]. Although the coupling of SAXS to Size-Exclusion Chromatography (SEC-SAXS) can, in some instances, separate the components of the mixture, there are multiple examples where the coexistence of multiple species is unavoidable. In these circumstances and with the aim to isolate the contribution of the individual species within complex mixtures, analytical approaches have been developed to decompose large SAXS titration datasets 9 [71,72]. This decomposition is easier when prior structural knowledge of the species is used for the analysis [70]. However, to apply this strategy to low-affinity flexible complexes, accurate conformational descriptions of all species in the free and bound forms are mandatory. The analysis of SAS data measured in samples with different relative concentrations of both partners seems the most appropriate strategy to enrich the information content in order to structurally characterize these extremely challenging scenarios (Figure 3). Conclusions and outlook During the last decade, SAS has been added to the toolbox of techniques used to study conformational fluctuations in proteins. This dynamic revolution of SAS is linked to the development of computational tools able to describe the conformational landscape of biomolecules and ensemble approaches with the capacity to interpret SAS data in terms of structural variability. These computational tools, which use chemical and structural knowledge of biomolecules, partially compensate for the limited amount of information coded in a SAS curve. Therefore, the capacity to fully exploit the structural information held in SAS data will necessarily be linked to the development of more advanced and precise computational approaches with specially developed force-fields. This notion is especially applicable to IDPs and IDRs, which populate a huge number of conformational states. For these proteins, SAS can be enriched with complementary information obtained by NMR, smFRET, EPR, or MS, and integrated into a common ensemble model embedding structure and dynamics. A particularly challenging subclass of IDPs is that containing Low-Complexity Regions (LCRs), which are involved in multitude of biological processes and are related to severe pathologies. LCRs are unusually simple protein sequences with a strong amino acid composition bias. The resulting similarity of chemical environments within their sequence hampers their structural characterization by NMR. SAS can be a valuable alternative through which to study this important but structurally neglected family of proteins [73,74,75]. The function of multitude of IDPs is determined by their interaction with biomolecular partners to form assemblies, which, in many cases, are of low to moderate affinity. The capacity of SAS to probe the size and shape of particles in solution places this technique in a unique position to address these polydisperse scenarios. A case in point is the 10 fibrillation process that several IDPs undergo to form amyloids, which are linked to severe diseases. The decomposition of time-dependent SAXS datasets has been successfully used to characterize intermediate oligomeric forms [76*,77], thereby validating SAXS as a practical tool for this purpose The need to understand the mechanisms underlying complex cellular processes and recent technical and conceptual advances in structural biology techniques across the board have prompted researchers to tackle challenging systems that were inaccessible some years ago. Many of these systems are inherently dynamic and/or polydisperse and can be exquisitely probed by SAS. As a consequence, we anticipate that SAS will take on greater relevance in hybrid approaches where its unique information will be synergistically integrated with data from multiple sources to deliver accurate structural and dynamic models of disordered proteins and their complexes. Conflict of interest statement Authors declare no conflict of interest Acknowledgements This work was supported by the ERC-CoG chemREPEAT, SPIN-HD – Chaires d'Excellence 2011 from the Agence National de Recherche (ANR), ATIP-Avenir, and the French Infrastructure for Integrated Structural Biology (FRISBI – ANR-10-INSB-05-01) to PB. FHT is supported by INSERM and the Sapere Aude Programme SAFIR of the University of Copenhagen. AU is supported by a grant from the Fondation pour la Recherche Médicale. 11 References and recommended reading [1] Dunker AK, Brown CJ, Lawson JD, Iakoucheva LM, Obradovic Z: Intrinsic disorder and protein function. Biochemistry 2002, 41:6573-6582. [2] Wright PE, Dyson HJ: Intrinsically disordered proteins in cellular signalling and regulation. Nat Rev Mol Cell Biol 2015, 16 :18-29. [3**] Csizmok V, Follis AV, Kriwacki RW, Forman-Kay JD : Dynamic Protein Interaction Networks and New Structural Paradigms in Signaling. Chem Rev 2016, 116:6424- 6462. Excellent review with a very complete list of references on the unique mechanisms used by disordered proteins to perform very specific functions in signaling processes. A description is provided of the present knowledge on the emerging field of phase separation induced by the interaction of IDRs with RNA. [4] Xie H, Vucetic S, Iakoucheva LM, Oldfield CJ, Dunker AK, Uversky VN, Obradovic Z : Functional anthology of intrinsic disorder. 1. Biological processes and functions of proteins with long disordered regions. J Proteome Res 2007, 6:1882-98 [5] Tompa P, Schad E, Tantos A, Kalmar L: Intrinsically disordered proteins: emerging interaction specialists. Curr Opin Struct Biol 2015, 35:49-59. [6] Dunker AK, Cortese MS, Romero P, Iakoucheva LM, Uversky VN: Flexible nets. The roles of intrinsic disorder in protein interaction networks. FEBS J 2005, 272:5129- 5148. [7] Kim PM, Sboner A, Xia Y, Gerstein M: The role of disorder in interaction networks: A structural analysis. Mol Systems Biol 2008, 4:179. [8] Dyson HJ, Wright PE: Unfolded proteins and protein folding studied by NMR. Chem Rev 2004, 104:3607-3622. [9] Jensen MR, Ruigrok RWH, Blackledge M: Describing intrinsically disordered proteins at atomic resolution by NMR. Curr Opin Struct Biol 2013, 23 :426-435. [10] Feigin LA, Svergun DI: Structure analysis by small-angle X-ray and neutron scattering. 1987, New York: Plenum Press. [11] Putnam CD, Hammel M, Hura GL, Tainer JA: X-ray solution scattering (SAXS) combined with accurate macromolecular structures, conformations and assemblies in solution. Quart Rev Biophys 2007, 40:191-285. computation: defining crystallography and [12] Jacques DA, Trewhella J: Small-angle scattering for structural for structural biology-expanding the frontier while avoiding the pitfalls. Protein Sci 2010, 19:642- 657. [13] Doniach S: Changes in biomolecular conformation seen by small angle X-ray scattering. Chem Rev 2001, 101:1763-1778. [14] Bernadó P, Svergun DI: Structural analysis of intrinsically disordered proteins by small-angle X-ray scattering. Mol Biosyst 2012, 8:151-167. 12 [15] Receveur-Brechot V, Durand D: How random are intrinsically disordered proteins? A small angle scattering perspective. Curr Protein Pept Sci 2012, 13:55-75. [16] Bernadó P, Blackledge M: Structural Biology: Proteins in dynamic equilibrium. Nature 2010, 468:1046-1048. [17] Zhou H-X: Polymer models of protein stability, folding, and interactions. Biochemistry 2004, 43:2141-2154. [18] Jha AK, Colubri A, Freed KF, Sosnick TR: Statistical coil model of the unfolded state: resolving the reconciliation problem. Proc Natl Acad Sci USA 2005, 102:13099- 13104. [19] Bernadó P, Blanchard L, Timmins P, Marion D, Ruigrok RW, Blackledge M : A structural model for unfolded proteins from residual dipolar couplings and small- angle x-ray scattering. Proc Natl Acad Sci USA 2005, 102:17002-17007. [20*] Ozenne V, Bauer F, Salmon L, Huang J, Jensen MR, Segard S, Bernadó P, Charavay C, Blackledge M: Flexible-meccano:a tool for the generation of explicit ensemble descriptions of their associated experimental observables. Bioinformatics 2012, 28:1463-1470. intrinsically disordered proteins and Description of Flexible-Meccano. This software computes ensembles of IDPs based on the conformational sampling found in coil regions of crystallographic structures. The program computes averaged RDCs and PREs from the ensembles, and provides scripts to add side chains, and to compute CSs and SAXS data. [21] Vitalis A, Pappu, RV: ABSINTH: A new continuum solvation model for simulations of polypeptides in aqueous solutions. J Comput Chem 2009, 30:673–699. [22] Best RB, Zheng W, Mittal J: Balanced protein-water interactions improve properties of disordered proteins and non-specific protein associantion. J Chem Theory Comput 2014, 10:5113−5124. [ 23 ] Henriques J, Cragnell C, Skepö M: Molecular dynamics simulations of intrinsically disordered proteins: Force field evaluation and comparison with experiment. J Chem Theory Comput 2015, 11:3420-3431. [24] Mercadante D, Milles S, Fuertes G, Svergun DI, Lemke EA, Gräter F: Kirkwood-Buff Approach Rescues Overcollapse of a Disordered Protein in Canonical Protein Force Fields. J Phys Chem B 2015, 119:7975-7984 [25] Chebaro Y, Ballard AJ, Chakraborty D, Wales DJ: Intrinsically disordered energy landscapes. Sci Rep 2015, 5:10386. [26] Zerze GH, Miller CM, Granata D, Mittal J: Free energy surface of an intrinsically disordered protein: Comparison between temperature replica exchange molecular dynamics and bias-exchange metadynamics. J Chem Theory Comput 2015, 11:2776-2782. [27] Lee KH, Chen J: Multiscale Enhanced Sampling of Intrinsically Disordered Protein Conformations. J Comput Chem2016, 37:550-557. [28] Dedmon M, Lindorff-Larsen K, Christodoulou J, Vendruscolo M, Dobson CM: Mapping long-range interactions in alpha-synuclein using spin-label NMR and ensemble molecular dynamics simulations. J Am Chem Soc 2005, 127:476-477. 13 [29]Mukrasch MD, Markwick P, Biernat J, Bergen Mv, Bernadó P, Griesinger C, Mandelkow E, Zweckstetter M, Blackledge M: Highly populated turn conformations in natively unfolded tau protein identified from residual dipolar couplings and molecular simulation. J Am Chem Soc 2007, 129:5235-5243. [30] Wu K-P, Weinstock DS, Narayanan C, Levy RM, Baum J: Structural reorganization of α-synuclein at low pH observed by NMR and REMD simulations. J Mol Biol 2009, 391:784-796. [31] Kohn JE, Millett IS, Jacob J, Zagrovic B, Dillon TM, Cingel N, Dothager RS, Seifert S, Thiyagarajan P, Sosnick TR, Hasan MZ, Pande VS, Ruczinski I, Doniach S, Plaxco KW: Random-coil behavior and the dimensions of chemically unfolded proteins. Proc Natl Acad Sci USA 2004, 101:12491-12496. [32] Bernadó P, Blackledge M: A self-consistent description of the conformational behavior of chemically denatured proteins from NMR and small angle scattering. Biophys J 2009, 97:2839-2845. [33] Mukrasch MD, Markwick P, Biernat J, Bergen Mv, Bernadó P, Griesinger C, Mandelkow E, Zweckstetter M, Blackledge M : Highly populated turn conformations in natively unfolded tau protein identified from residual dipolar couplings and molecular simulation. J Am Chem Soc. 2007, 129:5235-5243. [34] Zagrovic B, Lipfert J, Sorin EJ, Millett IS, van Gunsteren WF, Doniach S, Pande VS: Unusual compactness of a polyproline type II structure. Proc. Natl. Acad. Sci. USA 2005, 102:11698-11703. [35] Wells M, Tidow H, Rutherford TJ, Markwick P, Jensen MR, E. Mylonas, Svergun DI, Blackledge M, Fersht AR: Structure of tumor suppressor p53 and its intrinsically disordered N-terminal transactivation domain. Proc Natl Acad Sci USA 2008, 105:5762-5767. [36] De Biasio A, Ibáñez de Opakua A, Cordeiro TN, Villate M, Merino N, Sibille N, Lelli M, Diercks T, Bernadó P, Blanco FJ : p15PAF is an intrinsically disordered protein with nonrandom structural preferences at sites of interaction with other proteins. Biophys J 2014, 106:865-874. [37] Bernadó P, Mylonas E, Petoukhov MV, Blackledge M, Svergun DI : Structural characterization of flexible proteins using small-angle X-ray scattering. J Am Chem Soc 2007, 129:5656-5664. [38] Tria G, Mertens HD, Kachala M, Svergun DI : Advanced ensemble modelling of flexible macromolecules using X-ray solution scattering. IUCrJ 2015, 2:207-217. [39] Pelikan M, Hura GL, Hammel M: Structure and flexibility within proteins as identified through small angle X-ray scattering. Gen Physiol Biophys 2009, 28:174- 189. [40] Yang S, Blachowicz L, Makowski L, Roux B: Multidomain assembled states of Hck tyrosine kinase in solution. Proc Natl Acad Sci USA 2010, 107:15757-15762. [41] Bertini I, Giachetti A, Luchinat C, Parigi G, Petoukhov MV, Pierattelli R, Ravera E, Svergun DI : Conformational space of flexible biological macromolecules from average data. J Am Chem Soc 2010, 132:13553-13558. 14 [42] Rozycki B, Kim YC, Hummer G: SAXS ensemble refinement of ESCRT-III CHMP3 conformational transitions. Structure 2011, 19:109-116. [43] Daughdrill GW, Kashtanov S, Stancik A, Hill SE, Helms G, Muschol M, Receveur- Bréchot V, Ytreberg FM : Understanding the structural ensembles of a highly extended disordered protein. Mol Biosyst 2012, 8:308-319. [44] Antonov LD, Olsson S, Boomsma W, Hamelryck T : Bayesian inference of protein ensembles from SAXS data. Phys Chem Chem Phys 2016, 18:5832-5838. [45*] Shkumatov AV, Chinnathambi S, Mandelkow E, Svergun DI : Structural memory of natively unfolded tau protein detected by small-angle X-ray scattering. Proteins 2011, 79:2122-2131. Interesting article reporting on the conformational changes experienced by Tau protein when submitted to temperature jumps. Although the authors do not provide a precise explanation on the origin of this phenomenon, it reflects that there are probably structural features in IDPs that have not been unveiled yet. [46] Kjaergaard M, Nørholm AB, Hendus-Altenburger R, Pedersen SF, Poulsen FM, Kragelund BB : Temperature-dependent structural changes intrinsically disordered proteins: formation of alpha-helices or loss of polyproline II? Protein Sci 2010, 19:1555-1564. in [47] Leyrat C, Jensen MR, Ribeiro EA Jr, Gérard FC, Ruigrok RW, Blackledge M, Jamin M : The N(0)-binding region of the vesicular stomatitis virus phosphoprotein is globally disordered but contains transient α-helices. Protein Sci 2011, 20:542-556. [48] Stott K, Watson M, Howe FS, Grossmann JG, Thomas JO : Tail-mediated collapse of HMGB1 is dynamic and occurs via differential binding of the acidic tail to the A and B domains. J Mol Biol 2010, 403:706-722. [49] Sibille N, Bernadó P : Structural characterization of intrinsically disordered proteins by the combined use of NMR and SAXS. Biochem Soc Trans. 2012, 40:955- 962. [50] Jensen MR, Zweckstetter M, Huang JR, Blackledge M : Exploring free-energy landscapes of intrinsically disordered proteins at atomic resolution using NMR spectroscopy. Chem Rev 2014, 114:6632-6660. [51] Marsh JA, Neale C, Jack FE, Choy WY, Lee AY, Crowhurst KA, Forman-Kay JD: Improved structural characterizations of the drkN SH3 domain unfolded state suggest a compact ensemble with native-like and non-native structure. J Mol Biol 2007, 367:1494-1510. [52] Krzeminski M, Marsh JA, Neale C, Choy WY, Forman-Kay JD : Characterization of disordered proteins with ENSEMBLE. Bioinformatics 2013, 29:398-399. [53] Jensen MR, Houben K, Lescop E, Blanchard L, Ruigrok RW, Blackledge M : Quantitative conformational analysis of partially folded proteins from residual dipolar couplings: application to the molecular recognition element of Sendai virus nucleoprotein. J Am Chem Soc 2008, 130:8055-8061. [54**] Schwalbe M, Ozenne V, Bibow S, Jaremko M, Jaremko L, Gajda M, Jensen MR, Biernat J, Becker S, Mandelkow E, Zweckstetter M, Blackledge M: Predictive Atomic 15 Resolution Descriptions of Intrinsically Disordered hTau40 and α-Synuclein in Solution from NMR and Small Angle Scattering. Structure 2014, 22:238–249. Seminal study on the structural properties of α-synuclein and Tau. Using ASTEROIDS the authors interpreted complete CS, RDC and PRE datasets in combination with SAXS data to deliver conformational ensembles of both proteins. The most relevant part part of the article is the extensive cross-validation analyses that demonstrate the accuracy of the structural models. [55] Delaforge E, Milles S, Bouvignies G, Bouvier D, Boivin S, Salvi N, Maurin D, Martel A, Round A, Lemke EA, Jensen MR, Hart DJ, Blackledge M : Large-Scale Conformational Dynamics Control H5N1 Influenza Polymerase PB2 Binding to Importin α. J Am Chem Soc 2015, 137:15122-15134. [56] Boura E, Rózycki B, Herrick DZ, Chung HS, Vecer J, Eaton WA, Cafiso DS, Hummer G, Hurley JH : Solution structure of the ESCRT-I complex by small-angle X-ray scattering, EPR, and FRET spectroscopy. Proc Natl Acad Sci USA. 2011, 108:9437- 9442. [57**] Boura E, Różycki B, Chung HS, Herrick DZ, Canagarajah B, Cafiso DS, Eaton WA, Hummer G, Hurley JH : Solution structure of the ESCRT-I and -II supercomplex: implications for membrane budding and scission. Structure 2012, 20:874-886. In this study the ensemble description of the flexible complex formed by ESCRT-I and II is obtained. The authors integrate SAXS, smFRET and EPR data to define the complex. A protocol is introduced in order to find the minimal number of components in the ensemble with the capacity to properly describe the three sources of data. [58] Konermann L, Vahidi S, Sowole MA. Mass spectrometry methods for studying structure and dynamics of biological macromolecules. Anal Chem 2014, 86, 213-232. [59] Borysik AJ, Kovacs D, Guharoy M, Tompa P. Ensemble Methods Enable a New Definition for the Solution to Gas-Phase Transfer of Intrinsically Disordered Proteins. J Am Chem Soc 2015, 137, 13807-13817. [60] Keppel TR, Weis DD. Mapping residual structure in intrinsically disordered proteins at residue resolution using millisecond hydrogen/deuterium exchange and residue averaging. J Am Soc Mass Spectrom 2015, 26, 547-554. [61] O'Brien DP, Hernandez B, Durand D, Hourdel V, Sotomayor-Pérez AC, Vachette P, Ghomi M, Chamot-Rooke J, Ladant D, Brier S, Chenal A Structural models of intrinsically disordered and calcium-bound folded states of a protein adapted for secretion. Sci Rep 2015, 5, 14223. [62] Mylonas E, Hascher A, Bernadó P, Blackledge M, Mandelkow E, Svergun DI. Domain conformation of tau protein studied by solution small-angle X-ray scattering. Biochemistry 2008, 47:10345-10353. [63] Sharma R, Raduly Z, Miskei M, Fuxreiter M: Fuzzy complexes: Specific binding without complete folding. FEBS Lett 2015, 589: 2533–2542. [64] Shell SS, Putnam CD, Kolodner RD: The N terminus of Saccharomyces cerevisiae Msh6 is an unstructured tether to PCNA. Mol Cell 2007, 26:565–578. 16 [65*] Rochel N, Ciesielski F, Godet J, Moman E, Roessle M, Peluso-Iltis C, Moulin M, Haertlein M, Callow P, Mély Y, Svergun DI, Moras D: Common architecture of nuclear receptor heterodimers on DNA direct repeat elements with different spacings. Nat Struct Mol Biol 2011, 18:564–570. Using a combination of SAXS, SANS and smFRET the authors studied the structure of three hormonal nuclear receptor heterodimers in complex with cognate dsDNA and intrinsically disordered co-activators. Although using ab initio reconstructions, the asymmetric singly- bound nature of the complex with the co-activator is demonstrated. Excellent study that highlights the power of SAXS/SANS to characterize complex biomolecular entities. [66] Devarakonda S, Gupta K, Chalmers MJ, Hunt JF, Griffin PR, Van Duyne GD, Spiegelman BM: Disorder-to-order transition underlies the structural basis for the assembly of a transcriptionally active PGC-1α/ERRγ complex. Proc Natl Acad Sci USA 2011, 108:18678–18683. [67] Różycki B, Boura E : Large, dynamic, multi-protein complexes: a challenge for structural biology. J Phys Condens Matter. 2014, 26:463103. [68] De Biasio A, de Opakua AI, Mortuza GB, Molina R, Cordeiro TN, Castillo F, Villate M, Merino N, Delgado S, Gil-Cartón D, Luque I, Diercks T, Bernadó P, Montoya G, Blanco FJ : Structure of p15(PAF)-PCNA complex and implications for clamp sliding during DNA replication and repair. Nat Commun 2015, 6:6439. [69**] Yabukarski F, Leyrat C, Martinez N, Communie G, Ivanov I, Ribeiro EA Jr, Buisson M, Gerard FC, Bourhis JM, Jensen MR, Bernadó P, Blackledge M, Jamin M : Ensemble Structure of the Highly Flexible Complex Formed between Vesicular Stomatitis Virus Unassembled Nucleoprotein and its Phosphoprotein Chaperone. J Mol Biol 2016, 428:2671-94. In this study the ensemble structure of the viral complex between the nucleocapsid protein N0 and the phosphoprotein P is determined. The ensemble description is performed using the EOM approach. The novelty of the study is the simultaneous description of the SAXS data of the complex with SANS curves measured at four different contrast levels. This is the first study that profits from the rich information from contrast variation in the context of highly flexible biomolecular complexes using ensemble approaches. [70] Tuukkanen AT, Svergun DI : Weak protein-ligand interactions studied by small- angle X-ray scattering. FEBS J 2014, 281:1974-87. [71] Blobel J, Bernadó P, Svergun DI, Tauler R, Pons M : Low-resolution structures of transient protein-protein complexes using small-angle X-ray scattering. J Am Chem Soc 2009, 131:4378-86. [ 72 ] Chandola H, Williamson TE, Craig BA, Friedman AM, Bailey-Kellogg C : Stoichiometries and affinities of interacting proteins from concentration series of solution scattering data: decomposition by least squares and quadratic optimization. J Appl Crystallogr 2014, 47:899-914. [73] Greving I, Dicko C, Terry A, Callow P, Vollrath F : Small angle neutron scattering of native and reconstituted silk fibroin. Soft Matter 2010, 6:4389. 17 [74] Boze H, Marlin T, Durand D, Pérez J, Vernhet A, Canon F, Sarni-Manchado P, Cheynier V, Cabane B : Proline-rich salivary proteins have extended conformations. Biophys J 2010, 99:656–665. [75] Owens GE, New DM, West AP, Bjorkman PJ : Anti-PolyQ Antibodies Recognize a Short PolyQ Stretch in Both Normal and Mutant Huntingtin Exon 1. J Mol Biol 2015, 427:2507–2519. [76*] Vestergaard B, Groenning M, Roessle M, Kastrup JS, van de Weert M, Flink JM, Frokjaer S, Gajhede M, Svergun DI: A helical structural nucleus is the primary elongating unit of insulin amyloid fibrils. PLoS Biol 2007, 5:1089–1097. Pioneering study on the characterization of fibrillating proteins using SAXS. The fibrillation of insulin is monitored by SAXS in a time-dependent manner. The resulting curves are the population-weigted averages of all species co-existing in solution. In an arduous procedure, the species-pure curves for the three main components of the mixtures were decomposed allowing their structural characterization including their molecular weight, oligomerization state, and 3D arrangement. [77] Giehm L, Svergun DI, Otzen DE, Vestergaard B Low-resolution structure of a vesicle disrupting alpha-synuclein oligomer that accumulates during fibrillation. Proc Natl Acad Sci USA 2011, 108:3246–3251. 18 FIGURES: Figure 1. Rg values from 76 IDPs as a function of the number of residues of the protein are plotted in Log-Log scale. Only proteins lacking a permanent secondary or tertiary structure were considered for the compilation. Proteins with ordered domains, molten globules, or denatured proteins were not considered. Straight lines correspond to Flory's relationships parametrized for denatured proteins using experimental data (purple-dashed) [31] and IDPs using computational ensembles calculated with Flexible-Meccano (green-solid) [32]. Colored bands correspond to uncertainty of the parametrization for both models. Some IDPs contain local structural features and consequently they are globally more extended or more compact than expected for a random coil. These structural features, even if transient, can be manifested in the experimental Rg. 19 Figure 2. Cartoons representing the structural sensitivity of NMR, X-ray crystallography, and SAS for a complex involving a disordered protein (central cartoon). NMR normally probes the flexible regions of these complexes while the globular partner and the interacting region remain invisible. Crystallography provides detailed information of the interacting region of the complex but not for the flexible parts. SAXS probes the complete ensemble, although the details cannot be assessed due to its inherent low-resolution. SANS, through contrast variation experiments, can probe independently both partners in the context of the complex depending on the deuteration level of the partners and the D2O/H2O of the buffer. SAS is an ideal tool to integrate NMR and crystallographic information to build complete structural and dynamic models of disordered biomolecular complexes. 20 Figure 3. Examples of polydisperse scenarios that can occur in low-affinity complexes involving an IDP and a globular partner. (A) Both proteins have a single binding site. The complex is in equilibrium with the free forms of both proteins. (B) The globular partner is a dimer and has two identical binding sites. The free forms are in equilibrium with three possible complexes recognizing one or two binding sites of the globular partner. Due to the symmetry of the dimer, the two singly bound complexes are however indistinguishable by SAS. (C) The IDP presents two similar binding sites (pink and green). The free forms are in equilibrium with two 1:1 complexes using a distinct IDP interacting site to bind the globular partner, and a complex where the IDP simultaneously interacts with two globular partners. On the right part of the figure, three panels are displayed representing the molar fraction of each species along a simulated titration experiment for each scenario. These populations were computed assuming a fixed concentration of the globular partner, [globular] = 100 μM, and increasing concentrations of IDP, [IDP], from 1 μM to 400 μM. A common dissociation constant Kd = 20 μM was used for scenarios A and B, in panel C the two IDP binding sites, pink and green, display a Kd = 20 μM and 40 μM, respectively. These panels exemplify the inherent polydispersity moderate affinity complexes, and how multiple titration experiments will probe differently the species present and their relative populations. 21
1211.4788
2
1211
2015-11-16T20:00:02
Modal locking between vocal fold and vocal tract oscillations: Experiments and statistical analysis
[ "physics.bio-ph", "physics.class-ph" ]
The human vocal folds are known to interact with the vocal tract acoustics during voiced speech production; namely a nonlinear source-filter coupling has been observed both by using models and in \emph{in vivo} phonation. These phenomena are approached from two directions in this article. We first present a computational dynamical model of the speech apparatus that contains an explicit filter-source feedback mechanism from the vocal tract acoustics back to the vocal folds oscillations. The model was used to simulate vocal pitch glideswhere the trajectory was forced to cross the lowest vocal tract resonance, i.e., the lowest formant $F_1$. Similar patterns produced by human participants were then studied. Both the simulations and the experimental results reveal an effect when the glides cross the first formant (as may happen in \textipa{[i]}). Conversely, this effect is not observed if there is no formant within the glide range (as is the case in \textipa{[\textscripta]}). The experiments show smaller effect compared to the simulations, pointing to an active compensation mechanism.
physics.bio-ph
physics
Modal locking between vocal fold and vocal tract oscillations: experiments and statistical analysis D. Aalto1, J. Malinen2,3, and M. Vainio4 1 Speech Communication and Disorders, University of Alberta, Canada 2 Dept. Mathematics and Systems Analysis, Aalto University, Finland 3 Dept. Signal Processing and Acoustics, Aalto University, Finland 4 Inst. Behavioural Sciences (SigMe group), University of Helsinki, Finland [email protected] April 22, 2019 Abstract The human vocal folds are known to interact with the vocal tract acoustics during voiced speech production; namely a nonlinear source- filter coupling has been observed both by using models and in in vivo phonation. These phenomena are approached from two directions in this article. We first present a computational dynamical model of the speech apparatus that contains an explicit filter-source feedback mech- anism from the vocal tract acoustics back to the vocal folds oscillations. The model was used to simulate vocal pitch glides where the trajec- tory was forced to cross the lowest vocal tract resonance, i.e., the lowest formant F1. A more detailed analysis of the simulations is given in a companion article. Similar vocal glide patterns produced by human participants are studied in this article. Both the simulations and the experimental results reveal an effect when the glides cross the first formant (as may happen in [i]). Conversely, this effect is not observed if there is no formant within the glide range (as is the case in [A]). The experiments show smaller effect compared to the simulations, pointing to an active compensation mechanism. Keywords. Speech modelling, vocal folds model, flow induced vibrations, modal locking. PACS. Primary 4370Bk. Secondary 4370Gr, 4370Aj, 4370Mn. 1 1 Introduction It is an underlying assumption in the linear source -- filter theory of vowel production that the sound source (i.e., the vocal fold vibration) operates independently of the filter (i.e., the vocal tract, henceforth VT) whose res- onances modulate the resulting vowel sound [1, 2]. This classical approach captures a wide range of phenomena in speech production, at least in male speakers. However, significant observations remain unexplained by the feed- back free source-filter model, such as some fine structure in the phonation of a female singer near the lowest formant F1. Instability of the fundamental (glottal) frequency f0 has been detected acoustically [3] and at tissue level [4] when a singer performs an f0-glide over F1 on a steady vowel. As argued by Titze [5], the observed frequency jumps are due to a feedback coupling from the vocal sound pressure back to the glottal pulse generation, denoted there as the nonlinear source-filter coupling. Since F1 usually lies well above f0 in male phonation, this phenomenon occurs typically in female subjects when they are producing vowels with a low F1. In laboratory experiments, Titze et al. found more instabilities in male subjects, possibly because males have less experience in suppressing unwanted instabilities [3]. The vocal source instabilities have been modeled by low-order mass- spring systems. A two-mass vocal folds model, coupled with a resonator tube, showed coupling related effects when the dimensions of the tube were manipulated [6]. Tokuda et al. simulated vocal pitch glides using a four- mass model to analyze the interactions between vocal register transitions and VT resonances [7]. We give a brief treatment of our model simulations in this article; more detailed analysis using a more comprehensive phonation model is given in the companion article [8]. For the current study, the f0-F1 cross-over phenomenon is approached from two complementary directions: 1) by simulating f0-glides numerically on a steady vowel where F1 is near f0, using a dynamical vowel model that includes the filter-source feedback loop described above; and 2) by carrying out a related vowel glide production experiment on female test subjects. 2 The vocal folds model 2.1 Anatomy Before introducing the glottis model, some anatomic details are reviewed, and physiological control mechanisms (actuated by muscles) affecting the speech characteristics are explained. Such explanation should be regarded as an idealization since the effect of a single muscle contraction can to some extent be compensated by other muscles involved. The glottis model is based on a further simplification of this idealization. The vowel sound is produced by the self-sustained cut-off effect of the 2 (a) (b) Figure 1: (a): Sketch of the anatomy of the glottis according to Gray. (b): The geometry of the glottis model and the symbols used. The trachea (i.e. the channel leading from the lungs to glottis) is to the left in this sketch and the vocal tract is to the right. air flow from lungs, caused by a quasi-periodic closure of an aperture -- known as the rima glottidis -- by two string-like vocal folds. This process is called phonation, and the system comprising the vocal folds and the rima glottidis is known as the glottis. A single period of the sound pressure signal, produced by the glottal flow, is called the glottal pulse. As shown in Fig. 1a, each vocal fold consists of a vocal cord (also known as a vocal ligament) together with a medial part of the thyroarytenoid mus- cle, and the vocalis muscle (not specified in Fig. 1a). Both vocal folds are attached to the thyroid cartilage from their anterior ends and to two corresponding arytenoid cartilages (i.e., the left and the right) from their posterior ends; see Fig. 1a. In addition to these three cartilages, there is the ring-formed cricoid cartilage whose location is inferior to the thyroid car- tilage. The vocal folds and the associated muscles are supported by these cartilages as explained next. Between each of the arytenoid cartilages and the cricoid cartilage, there are two muscles attached. These are the posterior and the lateral cricoary- tenoid muscles, and they have opposite mechanical actions. Indeed, during phonation the vocal folds are adducted by the contraction of the lateral cricoarytenoid muscles, and conversely, abducted by the posterior cricoary- tenoid muscles during, e.g., breathing. This control action is realized by a rotational movement of the arytenoid cartilages in a transversal plane. In addition, there is a fifth (unpaired) muscle -- the arytenoid muscle -- whose contraction brings the arytenoid cartilages closer to each other, thus reducing the opening of the glottis independently of the lateral cricoary- tenoid muscles. These rather complicated control mechanisms regulate, in particular, the type of phonation in the breathy-pressed scale. 3 The fundamental frequency f0 of a vowel sound is controlled by an- other mechanism that is actuated by two cricothyroid muscles (not visible in Fig. 1a). The contraction of these muscles leads to a rotation of the thyroid cartilage with respect to the cricoid cartilage. Because of this rota- tion, the thyroid cartilage inclines to the anterior direction, thus extending the vocal folds. The elongation of the string-like vocal folds leads to in- creased stress which raises the fundamental frequency of their longitudinal vibrations. 2.2 Glottis model The anatomic configuration in Fig. 1a is modelled by a low-order mass-spring system shown in Fig. 1b, based on earlier work of Aalto [9, 11]. Numerical efficiency and theoretical tractability favor a low-order model, and the aim of the model is at functionality rather than at anatomic detail. The model is able to reproduce accurately the measured male glottal pulse obtained by inverse filtering [10, 12]. The model supports non-symmetric vocal fold vibrations, and both the fundamental frequency f0 as well as the phonation type can be chosen by parameter values. Register shifts (e.g., from modal register to falsetto) are important phenomena not in the scope of this model. Accounting for register shifts would require either modelling the vocal folds as aerodynamically loaded strings or by a high-order mass-spring system that has a string-like "elastic" behavior. The vocal fold model in Fig. 1b consists of two wedge-shaped vibrating elements that have two degrees of freedom each. The distributed mass of these elements can be reduced into three mass points which are located so that mj1 is at x = L, mj2 at x = 0, and mj3 at x = L/2. The elastic support of the vocal ligaments is approximated by two springs at points x = aL and x = bL. The equations of motion for the vocal folds are given by ( M1 W1(t) + B1 W1(t) + K1W1(t) = −F (t), M2 W2(t) + B2 W2(t) + K2W2(t) = F (t), t ∈ R. (1) where Wj = (wj1, wj2)T are the displacements of the right and left endpoints of the jth fold, j = 1, 2. The respective mass, damping, and stiffness matrices Mj, Bj, and Kj in (1) are 4 mj3 4 mj3 4 Mj =(cid:20) mj1 + mj3 Bj =(cid:20) bj1 and Kj = P (cid:20) a2kj1 + b2kj2 ab(kj1 + kj2) b2kj1 + a2kj2 (cid:21) . 4 (cid:21) , mj2 + mj3 ab(kj1 + kj2) 0 bj2 (cid:21) , 0 (2) The entries of these matrices are computed by means of Lagrangian mechan- ics. The damping matrices Bj are diagonal since the dampers are located 4 at the endpoints of the vocal folds. The springs are located symmetrically around the midpoint x = L/2, so that a = (L/2 + l)/L and b = (L/2 − l)/L. The control parameter P > 0 is used for simulating variable frequency f0- glides for the purpose of this work. During the glottal open phase (when ∆W1(t) > 0), the load terms of (1) are given by F = (FA,1, FA,2)T as given in Eq. (6). During the glottal closed phase (when ∆W1(t) < 0), there are no aerodynamic forces apart from the acoustic counter pressure from the VT, denoted by pc and properly introduced in the context of Eq. (7). Instead, there is a nonlinear spring force for the elastic collision of the vocal folds, given by the Hertz impact model [13]: F =" kH∆W13/2 − H0−H1/2 2L H1 2 h · pc H1 2 h · pc H0−H1/2 2L # . (3) The model geometry is shown in Fig. 1b, and it corresponds to the coro- nal section through the center of the vocal folds. As always in such biome- chanical modelling [7, 13, 16], the lumped parameters of the mass-spring system are in some correspondence to the masses, material parameters, and geometric characteristics of the sound producing tissues. Such parameters are, e.g., the mass and stiffness matrices Mi and Ki, i = 1, 2, that appear in the equations of motion for the vocal folds (1). More precisely, matrices Mi correspond to the vibrating masses of the vo- cal folds, including the vocal ligaments together with their covering mucous layers and (at least, partly) the supporting vocalis muscles. The elements of the matrices Kj are best understood as linear approximations of k(s) = f /s where f = f (s) is the contact force required for deflection s at the center of the string-like vocal ligament in Fig. 1a. It should be emphasized that the exact numerical correspondence of tissue parameters to lumped model parameters Mj and Kj is intractable (and for practical purposes, even ir- relevant), and their values in computer simulations must be tuned using measurement data of f0 and the measured form of the glottal pulse [10]. Even though the equations of motion are separate for both vocal folds, the parameters (hence, the simulated vocal folds movements) are symmetric in all of the simulations reported in this paper. 3 Full model of vowel production Having treated the modelling of the vocal folds, it remains to review the other components of the full model, i.e., the 1D incompressible flow model and the VT acoustics model. These two subsystems are coupled so that the flow depends on the time-dependent glottal opening. Conversely, the flow produces aerodynamic forces on the vocal folds, and it also acts as the acoustic source to the resonat- ing VT, modelled by Webster's equation. The acoustics of the sub-glottal 5 air cavities is not modelled at all. The VT sound (counter) pressure gives rise to the filter-source feedback from VT to the glottal oscillations. With- out this feedback, modal locking does not appear at all as can be verified by running model simulations with f0 ≈ F1 with pc = 0 in (6). 3.1 Flow An incompressible 1D flow through the glottal opening with velocity vo is described by vo(t) = 1 CinerhH1 (cid:18)psub − Cg ∆W1(t)3 vo(t)(cid:19) (4) where the latter term inside the parentheses (representing the viscous pres- sure loss) is motivated by the Hagen -- Poiseuille law in a narrow aperture. The constant sub-glottal pressure (subtracted by the ambient air pressure) is denoted by psub, and h is the width of the flow channel that is assumed rect- angular. The parameter Ciner regulates the flow inertia, and Cg regulates the pressure loss in the glottis. Aalto et al. observed that Ciner effectively re- veals the phonation type when other model parameters are estimated based on recorded speech signals [10]. The viscous pressure loss in (4) depends on the glottal opening at the narrowest point, i.e., ∆W1. At the other end (i.e., towards the trachea) the opening is ∆W2. These are given by (1) through ∆W2 (cid:21) = W2 − W1 +(cid:20) g (cid:20) ∆W1 H0 (cid:21) . (5) The parameter g is the glottal opening when there is no flow and the vocal folds do not vibrate (W1 = W2 = 0). In human anatomy, the parameter g is related to the position and orientation of the arytenoid cartilages. In the glottis, the flow velocity V (x, t) is assumed to satisfy the mass conservation law H(x, t)V (x, t) = H1vo(t) for incompressible flow where H(x, t) is the height of the flow channel inside the glottis. In the model geometry of Fig. 1b, we have H(x, t) = ∆W2(t) + x L (∆W1(t) − ∆W2(t)), x ∈ [0, L]. Now the pressure p(x, t) in the glottis is given by the two equations above and the Bernoulli law p(x, t) + 1 2 ρV (x, t)2 = psub for static flow. Since each vocal fold has two degrees of freedom, this pressure and the VT counter pressure pc can be reduced to a force pair (FA,1, FA,2)T where FA,1 affects at the narrow (superior) end of the glottis (x = L) and FA,2 at the wide (resp. inferior) end (x = 0). This reduction is carried out by using the total force and moment balance equations FA,1 + FA,2 = hZ L 0 6 (p(x, t) − psub) dx and L · FA,1 = hZ L 0 x(p(x, t) − psub) dx − pc · h H1 2 H0 − H1/2 2 . The moment is evaluated with respect to point (x, y) = (0, 0) for the lower fold and (x, y) = (0, H0) for the upper fold in Fig. 1b. Evaluation of these integrals yields FA,1 = 1 2 ρv2 FA,2 = 1 2 ρv2 ohL(cid:16)− ohL(cid:16)   H 2 ∆W1(∆W2−∆W1) + 1 H 2 ∆W2(∆W2−∆W1) − 1 1 H 2 (∆W1−∆W2)2 ln(cid:16) ∆W2 (∆W1−∆W2)2 ln(cid:16) ∆W2 ∆W1(cid:17)(cid:17) − H1(H0−H1/2) ∆W1(cid:17)(cid:17) + H1(H0−H1/2) (6) H 2 4L 4L 1 hpc, hpc. 3.2 Vocal tract The vocal tract acoustics is modeled by Webster's lossless horn resonator. The governing equation for the velocity potential Ψ(s, t) is Ψtt(s, t) − c2 A(s) ∂ ∂s(cid:18)A(s) ∂Ψ(s, t) ∂s (cid:19) = 0 (7) where c is the sound velocity. The parameter s ∈ [0, LV T ] is the distance from the narrow (superior) end of the glottis measured along the VT center line, and LV T is the length of the VT. The area function A(·) is the cross- sectional area of the VT, perpendicular to the VT center line. The sound pressure is given in terms of the velocity potential by p = ρΨt. At the glottis end, the resonator is controlled by the flow velocity vo from Eq. (4) through the boundary condition Ψs(0, t) = −vo(t). The resonator exerts a counter pressure pc(t) = ρΨt(0, t) to the vocal folds equations (1) through Eqs. (6), thereby forming a filter-source feedback loop. The bound- ary condition at lips is a frequency-independent acoustic resistance of the form Ψt(LV T , t) + θcΨs(LV T , t) = 0 where θ is the normalized acoustic re- sistance [14]. 3.3 Model parameters The glottal flow equations (4) contain two independent parameters psub/Cg and Ciner; the mass-spring system contains three parameter matrices Mj, Bj, and Kj for j = 1, 2; and the resonator equations contain the area func- tion A(·) in (7) and the acoustic termination parameter θ at the mouth. Out of these parameters, psub/Cg and θ are determined from physical considerations and A(·) from anatomical data obtained by MRI. The area function used in simulations corresponds to [ø:] as in Hannukainen et al [15]. The parameter range for Ciner has then been determined and validated so as to produce a realistic time-domain glottal pulse form [10]; the values corresponding to pressed phonation are used in this work. All these model 7 parameter values introduced so far are equally valid for both female and male phonation. It remains to consider the parameter matrices in the vocal folds equations (1) where the differences between female and male phonation are significant. Hor´acek et al. provide parameter values in male phonation [16, 13]. The parameter values (K1, K2) have been estimated indirectly by requiring the correct fundamental frequency f0 [9]. The focus of this paper is in female phonation, and it is difficult to produce similar data for female subjects using literature. Thus, the typical nominal "male versions" of parameters Mj and Kj for j = 1, 2 (as given in Aalto [9, 10]) are scaled so as to obtain typical "female version" as explained above. This scaling is based on the data given by Titze [17]. Dimensional analysis and scaling yield Mfemale = λαMmale and Kfemale = λα−2Kmale where the exponent α ∈ [2, 3] by physical grounds. Based on experimenta- tion, the value used here is α = 2.3. The scale factor λ ≈ 0.6 is suggested by Titze [17] but here the value λ = 0.55 produces more female like phonation in simulations. The resulting increase in the resonances of (1) by factor 1/λ ≈ 1.8 reflects correctly the higher pitch of the female voice. The f0-glide is simulated by additional scaling of the matrices Kj whereas the matrices Mj are kept constant. This is based on the assumption that the vibrating mass of vocal folds is not significantly reduced when the speaker's pitch increases; a reasonable assumption as far as register changes are ex- cluded. The authors would like to remark that the relative magnitudes of Mj and Kj essentially determine the resonance frequencies of model (1). However, attention must be paid to their absolute magnitudes using, e.g., dimensional analysis since otherwise the load terms ±F (t) in (1) (contain- ing the aerodynamic forces, contact force between the vocal folds during the glottal closed phase, and the counter pressure from the VT) would scale in an unrealistic manner. The damping parameters bji, i, j = 1, 2, in Eq. (2) play an important but problematic role in glottis models. If there is too much damping (while keeping all other model parameters fixed), sustained oscillations do not oc- cur. Conversely, too low damping will cause instability in simulated vocal folds oscillations. The magnitude of physically realistic damping in vibrating tissues is not available, and the present model could possibly fail to give a quasi-stationary glottis signal even if realistic experimental damping values were available for use. For simplicity, we set bji = β > 0 for i, j = 1, 2, and the value of glottis loss β is adjusted separately for each parameter value set in order to obtain stable but sustained oscillation. In particular, param- eter β is adjusted every 100 ms during the f0-glide simulations presented below. The damping remains always so small that its lowering effect on the resonances of the mass-spring system (1) is negligible. 8 Let us conclude by discussing the parameter magnitudes in Eq. (1). The total vibrating mass is mj1+mj2+mj3 = 0.48 g for male and 0.12 g for female phonation. The total spring coefficients are kj1 + kj2 = 193 N/m for male and 161.3 N/m for female phonation using the nominal values, i.e., when P = 1 in Eq. (2). The nominal values yield f0 = 110 Hz for male and 187 Hz for female phonation. If the characteristic thickness of the vocal folds is assumed to be about 5 mm, these parameters yield a magnitude estimate for the elastic modulus of the vocal folds by E ≈ kj1+kj2 · 5 · 10−3 m ≈ 6.6 kP a. This is in good comparison with Fig. 7 in Chhetri et al.[18] where estimates are given for the elastic modulus of ex vivo male vocal folds between 2.0 kP a and 7.5 kP a for different parts of the vocal folds. Lh 3.4 Numerical realization The model equations are solved numerically using MATLAB software and custom-made code. The vocal fold equations of motion (1) are solved by the fourth order Runge -- Kutta time discretization scheme. The discontinuity of the load F (t) at ∆W1(t) = 0 is dealt with by an interpolation procedure detailed in Aalto (Section 2.4)[9]. The flow equation (4) is solved by the backward Euler method. The VT is discretized by the FEM using piece- wise linear elements (N = 100) and the physical energy norm of Webster's equation. Crank -- Nicolson time discretization is used, and the time step is always 20 µs. 4 Simulation results Frequency glides of vowel [ø:] are simulated near the lowest formant F1 or its subharmonic F1/2. In these simulations, F1 (determined from spectrograms of simulated vowel signals) coincides with the lowest resonance of the VT (solved independently from the eigenvalue problem associated to (7) and the boundary conditions). To produce the glides, parameter P in Eq. (2) is in- creased (or decreased) as a quadratic function of time. Then the oscillation frequency of the vocal fold model (1) in the absence of the counter pressure pc -- denoted by f0 -- increases (respectively, decreases) as a linear func- tion of time. This can be observed in simulation results if the filter-source feedback mechanism is disconnected by setting the counter pressure pc = 0 in Eq. (6). When using pc(t) = ρΨt(0, t) with Eq. (7), the observed funda- mental frequency f0 of the whole system does not behave linearly in time (in contrast to f0) but exhibits jumps near F1 (and F1/2) as shown in Fig. 3b. This behavior, the modal locking, is due to the filter-source feedback [5, 4]. Changing the area function A(·) in Eq. (7) to correspond some other vowel than [ø:] does not change the results of model simulations apart from formant, and hence, modal locking frequencies. 9 4.1 f0 -- F1 crossover Frequency f0-glides of length 2 s are simulated by varying the parameter P of the glottis model so that 350 Hz < f0 < 810 Hz linearly in time. The increasing phase is shown in the spectrogram Fig. 3a with an auxiliary line showing the glide of f0 and another line showing F1 = 647 Hz. It is observed that f0 coincides first with f0, but then it suddenly jumps upwards to F1 when it reaches about 470 Hz. The wave form of the glottal pulse near the transition is a superposition of two signals with frequencies f0 and F1. When f0 exceeds F1, then f0 and f0 coincide again. The qualitative behavior is sketched in the rising part of Fig. 3b. In a downwards glide the behavior is almost symmetric. First f0 = f0 descends down to F1 where f0 locks for a while. The latent glottal model frequency f0, of course, goes down linearly without change. After a time lag, f0 is released from F1 and drops suddenly to f0. This behavior is shown in the descending part of Fig. 3b. As can be seen in Fig. 3b, there are two periods (of length Tu, Td > 0) during which f0 = F1. Let us denote the corresponding frequency jumps by φu, φd > 0. By Fig. 3b we see φu/Tu = φd/Td is determined by the ascend and descend rate of the simulated, symmetric f0-glide. We observe φd > φu consistently in all simulations, and if the f0-glide is very slow, we observe that φd ≈ φu and Fig. 3b becomes symmetric. Simulating extremely slow glides (during which the parameter P can be regarded as a constant), it is observed that the jumps φd and φu have a common lower bound φ ≈ 174 Hz. This number can be regarded as the "true" magnitude of the simulated frequency jump, and it is in reasonable correspondence with the results reported in [3, 5]. It was noted above that φd > φu. One way to understand this asymmetry is in terms of the energy dissipation from VT resonance modes. The only energy losses in Eq. (7) are due to the dissipative boundary condition at lips. During a downwards glide, there is a locking of f0 at F1, and then a lot of energy is contained in the corresponding eigenmode of the joint system, comprising the VT air column and the vocal fold masses. For the vocal folds oscillations to get "unlocked" from this eigenmode, most of this energy must first be dissipated either by mouth radiation or by dispersion to other eigenmodes as a consequence of nonlinearity and lack of time invariance of the model. The dissipation rate may be slow compared to the speed of decrease of the f0-glide. Thus, the latent frequency f0 may have fallen far below F1 before unlocking of f0 may take place. It is therefore expected that the time delay Td is sensitive to relative magnitudes of losses and amounts of energy stored in vocal folds and VT, and this is supported by the simulations. Indeed, Td becomes very large if the vocal folds oscillation amplitude is small during the glide. 10 4 3.5 3 2.5 2 1.5 1 0.5 ) z H k ( y c n e u q e r F 0 0.1 0.2 0.3 0.4 0.5 Time (s) 0.6 0.7 0.8 0.9 Figure 2: (a): Upwards f0-glide over F1/2. (b): Downwards f0-glide over F1/2. 4.2 Glides near subharmonics of F1 Similar f0-glides are produced as explained above but now 187 Hz < f0 < 390 Hz linearly during 2 s time interval. The subharmonic F1/2 = 324 Hz lies in this interval, and two kinds of behavior are observed. First, f0 increases from below in Fig. 2a until F1/2 is reached; there is no locking at F1/2 but f0 jumps one octave up and locks at F1. Second, the phenomenon does not appear at all if the vocal fold oscillation amplitude is large enough. In Fig. 2b there is a downward glide near F1/2. First and third harmon- ics of f0 have a static part matching F1 and 2F1 respectively. It may be possible that f0 is "partly locked" to F1/2 but there is a clear component that develops with f0. Such a complicated spectral behavior emphasizes the surprising dynamical richness of the vowel production model presented in this paper. 5 Glide production experiment The simulation results and earlier experimental evidence [3] suggest that consequences of filter-source feedback should be detectable at crossings of frequencies f0 and F1. Unfortunately, determining the precise crossing time of f0 and F1 in vowel glides is difficult when based on the acoustic signal alone. Indeed, the harmonics 2f0, 3f0, . . . do not coincide with the band- width of F1 if f0 ≈ F1. If f0 follows successfully a slowly varying reference glide (as should happen in these experiments), it is then far from a per- sistently exciting signal that would be required for precise detection of F1. In a similar glide production experiment, Titze et al. [3] asked the test subjects to produce a spectrally rich vocal fry adjacent to every glide. A good estimate for F1 can be obtained this way, but the formant may creep during the glide. The subject is, in fact, expected to use various techniques to avoid modal locking in order to produce audibly clean vowel glides, and F1 position is expected to be affected near f0. We conclude that obtain- 11 4 3.5 3 2.5 2 1.5 1 0.5 ) z H k ( y c n e u q e r F F = 647 Hz 1 T u φ u T d φ d 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0.2 0.4 0.6 0.8 1 Time (s) 1.2 1.4 1.6 1.8 0 0 0.5 1 1.5 2 Time (s) 2.5 3 3.5 4 Figure 3: (a): f0-glide 350 Hz -- 810 Hz. (b): A sketch of the modal locking in an f0-glide over F1 first upwards and then downwards. The thick (thin) line shows f0 (resp. f0) during the glide. ing the crossing time of f0 and F1 from acoustic signals remains a difficult estimation problem. Rather than replicating the experimental arrangements of Titze et al. [3], a complementary approach is taken here without F1 as a controlled variable. This is achieved by eliciting vocal glides for two different vowels [A] and [i], where [i] has F1 within the f0 glide range, and F1 of [A] is as far as possible from the same range. The model simulations predict that the f0-trajectories corresponding to [A] and [i] should have distinctly different characteristics that should be detectable in natural glide productions. 5.1 Quantification of the jumping pattern During a modal locking episode, f0 is expected to jump as in Fig. 3. To quantify the jump effect for statistical analysis, the Duration Ratio (DR) of the glide f0 = f0(t) is defined. For frequencies fA < fa < fb < fB, DR is defined by DR(f0) = tb − ta tB − tA (8) where (tA, tB) is the longest open interval where fA ≤ f0(t) ≤ fB for all t ∈ (tA, tB), and (ta, tb) is the longest open interval where fa ≤ f0(t) ≤ fb for all t ∈ (ta, tb) such that [ta, tb] ⊂ (tA, tB), i.e., tA < ta ≤ tb < tB. This implies 0 ≤ DR(f0) ≤ 1 for any glide f0. By definition, DR does not depend on the direction or the speed of the glide in the sense that DR(f0) = DR( f0) = DR(f a 0 (t) = f0(at) for a > 0. 0 ) where f0(t) = f0(−t) and f a In the current experiments, all octave glides are between 200 Hz and 400 Hz, and the parameter values for Eq. (8) are always fA = γ · 200 Hz ≈ 230 Hz, fa = γfA ≈ 264 Hz, fb = γfa ≈ 303 Hz, and fB = γfb ≈ 348 Hz where γ = 21/5 ≈ 1.148698. Hence, the octave [200 Hz, 400 Hz] is divided into five intervals equal in logarithmic scale. We say that the glide f0 is full if its 12 range contains all of the interval [fA, fB]. If f0(t) = a2kt for some a > 0, k 6= 0, and t ∈ [0, T ] is a full glide, then we call it logarithmically linear and we have DR(f0) = 1/3 which can be regarded as the nominal value of DR on full glides in the absence of any perturbations. For an ideal logarithmically linear full glide f0 with a single modal locking "jump" as in Fig. 3, we have DR(f0) < 1/3 if the jump intersects fb but not fa or fB. Similarly, DR(f0) > 1/3 if the jump intersects fa but not fb or fA. The jump size and the jump position F (satisfying F = F1 in glide simulations) change the DR in a practically convenient way: to see this, a numerical experiment was performed. Full, logarithmically linear jump patterns were constructed using jump size of 100 Hz and one million tokens of F drawn from a normal distribution with µ = 300 Hz and σ = 50 Hz. This produced E[DR] = 0.308 and SD[DR] = 0.325. Keeping σ constant but varying µ gave the following values: E[DR] = 0.354 and SD[DR] = 0.270 at µ = 250 Hz; E[DR] = 0.240 and SD[DR] = 0.270 at µ = 350 Hz; and finally E[DR] = 0.333 and SD[DR] = 0.002 at µ = 600 Hz as expected. The experiment is designed so as to create situations where glides of [i] are more likely to have lower DR compared to glides of [A], due to modal locking induced f0-jumps over fb for [i] but not at all for [A]. More precisely, the position of F1 of [i] is expected to be roughly at 300 Hz (which, of course, explains the choice of fb above), and the position of F1 of [A] is likely to be above 600 Hz. Thus, simulated model predictions of jump patterns can be formulated as the statistically testable hypotheses: Hypothesis 1. The population mean of the Duration Ratio on full [i]-glides f [i] 0 and on full [A]-glides f [A] 0 satisfies EhDR(f [i] 0 )i < EhDR(f [A] 0 )i. Hypothesis 2. The population mean of the Duration Ratio on full [i]-glides f [i] 0 Ratio on full [A]-glides f [A] 0 0 )i < 1/3, and the population mean of the Duration satisfies EhDR(f [i] satisfies EhDR(f [A] 0 )i = 1/3. 5.2 Subjects and the recording arrangement Eleven native Finnish speaking female students at the University of Helsinki were recruited for the experiment. Females were chosen as subjects because (non-singing) males have typically little familiarity of using their vocal range around their naturally occurring F1. Hence, using males would probably lead to an overly high rejection rate of glide samples and excursions to falsetto register in otherwise successful glides. None of the participants reported any hearing or voice problems, and they had not received professional training in singing although some of them had a musical hobby (like violin playing). The participants were informed of the general purpose of the research and presented with the gliding task. There was a short familiarization period before the actual experiment during 13 which the participants could practice the gliding with the vowel [A]. Some participants were helped to find the right pitch range. The data was recorded in a sound-proof studio with a high-quality microphone (AKG C4000B), and digitized with Digidesign (DIGI 002) and ProTools v.9. 5.3 Glide imitation stimuli The subjects were instructed to produce frequency glides by following a spectrally neutral, synthesized reference glides from 200 Hz to 400 Hz. Every reference glide started and ended with a constant frequency part of 0.25 s to create a sensation of the right initial (resp., final) pitch of the glide. The instruction was given as a pre-recorded 0.5 s sample of either of the vowels, followed by three beeps (countdown beep: 0.25 s signal followed by 0.75 s silence). The beeps had the pitch of the onset value of the desired glide (either 200 Hz or 400 Hz). The glides and beeps were frequency modulated triangle waves; the triangle wave function s(t) is a 2π-periodic function with constant slope (except for multiplicities of 2π). The upward glides are given by s(2πω(t)) with ω(t) = T ·200Hz (2t/T − 1) for t ∈ [0, T ] where T is the duration of the glide (either 1.5 s or 3.0 s). The instantaneous frequency of such a glide is given by ∂ω(t) ∂t = 2t/T · 200Hz which was assumed to be equal with the perceived pitch. log 2 5.4 Experimental setup The subjects were asked to imitate the pitch of reference glides by producing them with a prolonged, spoken [A] or [i]. The subjects heard the reference glide from earphones at a volume that was adjusted for each subject to be as loud as possible without being unpleasant. The purpose of this arrangement was to hinder the subjects from hearing their own production easily and to elicit good phonation. The data was gathered following a factorial 23 design with direction (either fall or rise), duration (either fast 1.5 s or slow 3 s), and vowel (either [A] or [i]) as factors. Each subject had her own pseudo-randomized list of stimuli. The lists were constructed so that the different stimulus types would be randomly and evenly distributed in the trial: the stimuli were divided in blocks of 24 where each stimulus condition occurred three times, three identical stimuli were never in a row (not even over the blocks), four identical glide directions, vowels, or glide speeds were never in a row. 5.5 Data analysis Three subjects were disqualified since they did not learn the gliding pattern well enough. Most of their productions had at least one of the following characteristics: the glide was one octave away from the guiding glide signal, 14 400 348 303 264 230 200 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 Figure 4: A typical production of a vowel glide [i]. The linear line depicts the guiding signal and the zig-zagging line shows the fundamental frequency of the glide. The intervals corresponding to the mid-frequency and total durations are shown as horizontal lines. the glide pattern was missing (static or undulating pitch), or the range of the glide was limited to less than a quarter. The remaining eight subjects had no difficulties in the gliding task. A total of 864 glides were processed further. The vocal pitch was ex- tracted by cross-correlation using Praat (Praat 5.2.21; default parameters with admissible range from 200 Hz to 400 Hz). The pitch trajectories of the glides were automatically chopped from the audio signal, based on the tim- ing of the guiding signals and further analyzed in MATLAB. A custom-made algorithm was used to calculate DR(·) from the data. The glides were analyzed for their "fullness" to exclude the related bias in Duration Ratio values. The range from fA to fB was divided in ten equally long frequency bands in logarithmic scale. If some of the bands did not contain a value of f0, the glide was not considered full. This led to exclusion of 8 glides (1%). A mixed linear model was fitted to the remaining data using statistical software package R (version 2.14.0) and the model parameter selection was based on log-likelihoods [19]. Both of the Hypotheses 1 and 2 were independently tested by t-tests. A typical production is shown in Fig. 4 where the f0 trajectory of a fast falling [i] is given. The reference glide is the diagonal line with flat ends at t = 0, 2. The four horizontal lines indicate the critical frequencies fA, fa, fb and fB as introduced above. The corresponding time intervals for Eq. (8) are given by (tA, tB) = (0.52, 1.59) and [ta, tb] = (0.77, 1.03), yielding DR(f0) = 0.243 < 1/3. 15 5.6 Results The estimated means and standard deviations for DR at each eight factor combination groups are shown in the Table 1. The estimated mean DR's of [A] are consistently larger than those of [i] (see Hypothesis 1), and also the nominal DR value 1/3 = 0.33 . . . is above the estimated mean of DR for all four [i] conditions. Table 1: The means and standard deviations for the Duration Ratio (DR) over each factor combination. Condition Mean Fast fall [A] 0.341 Fast fall [i] 0.324 Fast rise [A] 0.331 Fast rise [i] 0.312 Slow fall [A] 0.337 Slow fall [i] 0.326 Slow rise [A] 0.340 Slow rise [i] 0.320 SD 0.099 0.108 0.084 0.073 0.106 0.104 0.102 0.105 N 106 105 108 108 107 106 108 108 Considering the four vowel pair samples separately, Hypothesis 1 on the population mean DR holds at p ≤ 0.05 by fast rises only (Welch two sample (one-sided) t-test with df = 209, t = −1.8, p = 0.04). The estimated mean and standard deviation of DR using all [i] glides are 0.320 and 0.098 (n = 427), respectively. For all [A] glides, the respective values are 0.337 and 0.098 (n = 429). Hence, Hypothesis 1 is verified using the full data set (Welch two sample test with df = 853, t = −2.5, p < 0.01). The first part of Hypothesis 2 is verified at p = 0.05 by fast rise [i] (one- sided t-test, df = 107, t = −3.1, p = 0.001) and also by using all [i] glides as the sample (one-sided t-test with df = 426, t = −2.7, p < 0.01). The population mean of DR over all [A] is not significantly different from the nominal value (df = 428, t < 0.9, p < 0.4) but the probability of a Type II Error is still large. Thus, the latter part of Hypothesis 2 is not supported statistically conclusively by the current dataset. A mixed effects model supports the previous observations. Glide j = 1, . . . , ni from subject i produces the DR value denoted by yi,j, given by the mixed effects model yi,j = β0 + di,jβ1 + si,jβ2 + vi,jβ3 + bi + ei,j where the fixed factors (with numerical levels 0 or 1) are the glide direction di,j, the glide speed si,j, and the vowel type vi,j. The test subject related random factors bi ∼ N (µi, σi) are independent from the experimental errors ei,j ∼ N (0, σ). As can be seen in the Table 2, the vowel type has a significant 16 Table 2: The linear mixed-effects model results for Duration Ratio (DR). The fixed factors are vowel, direction and speed; the subjects are treated as random factors. Factor Intercept Vowel ([i]) Direction (rise) Speed (slow) Estimate Std. Error t-value p-value 0.338 -0.017 -0.007 0.004 0.009 0.007 0.007 0.007 39.0 -2.5 -0.98 0.56 0 0.01 0.3 0.6 effect (ANOVA, df = 845, t > 2.5, p = 0.011) while the effect of the direction and speed are not significant (df = 845, t < 1). In a more complicated model taking into account all the interactions of the factors, the log-likelihoods of the models were lower, and the interactions were never significant. 6 Discussion As is usual in model validation experiments, the measurement data is much more ambiguous than the results of numerical simulations. The reconcilia- tion of simulated and measured data is difficult because of the following two main challenges: 1. The computational model could give an unrealistically strong indica- tion of modal locking because of unmodelled physics (such as subglot- tal acoustics, supraglottal turbulence, and energy losses). 2. The humans have active control mechanisms (without counterparts in the computational model) that are expected to reduce the observable effects of modal locking. Subglottal acoustics has been completely excluded from the computational model. The main reason for this is the lack of experimental data that would be required to reproduce the extremely difficult absorbing boundary condi- tion that the progressively subdividing system of bronchi and the alveoles constitute. Moreover, in the four-mass glottis model of Tokuda et al. [7], subglottal acoustics play a minor role compared to the supraglottal acous- tics. Hence, the coupling of the subglottal acoustics and the vocal fold oscil- lations does not seem to be the primary source of the discrepancy between simulations and experimental results. The supraglottal aerodynamics might have influence on the vowel glides. State-of-the-art Computational Fluid Dynamics (CFD) models indicate sig- nificant vorticity above the glottis [20]. The vorticity-induced aerodynamic force to the vocal folds is a genuinely fluid mechanical feature that is beyond 17 the scope of the acoustic theory of speech as well as simple Bernoulli flow models in VT. Evaluating the contribution of this force would also require a more refined CFD model than is usually available. In low frequency glide productions, the larynx moves vertically [21]. However, the simulated vowel glides were produced using a fixed area func- tion, and no movement in the larynx area is taken into account at all. The increased pitch is a consequence of the elongation and higher stress in vocal folds due to rotation of the thyroid cartilage in the anterior direction. To- wards the end of a glide production, larynx moves as a consequence of both the contracting thorax and the control actions required for variable pitch. These two mechanisms either oppose or reinforce each other depending on the direction of the glide. All this applies equally to VT geometries of [A] and [i]. Considering the latter challenge 2 above, the direct modeling of the audi- tory, motor, and muscular mechanisms behind the task performance seems unfeasible at the moment. Fortunately, the experiments can be arranged in a manner that downplays or even interferes with the active control mech- anisms. For this reason, trained singers are -- somewhat paradoxically -- excluded as test subjects even though producing vowel glides resembles very much a singing exercise. (Recall that non-singing males were excluded for the opposite reason.) The drawback of having non-singers was that some of the subjects could not perform the glide task satisfactorily which may have led to selection bias. However, these subjects are expected to be less familiar in their vocal pitch control which would lead to less accurate com- pensation (and hence, better detection in experiments) of the modal locking disturbance. Control mechanisms based on auditory feedback have been shown to rapidly and accurately compensate deviations from target glides in glide production [22]. This could lead to asymmetric f0 trajectories in terms of the glide direction because of different delays in control and observation. Recalling the simulated f0 trajectory in Fig. 3b, such a modal locking event during a rising glide is detectable only right after a large jump has already occurred while a modal-locking event during a fall can, in principle, be detected when the f0 has just ceased to decline. Considering the control delay alone, suppose that the compensation strategy against modal locking were only based on the auditory observation with an identical delay (of say 80 ms) for both rising and falling glides. Then the fall would be less perturbed by the modal locking event because then the control action has more time to react. For obvious reasons, however, the auditive observation of modal locking events in falling glides could be more delayed compared to rising glides for f0 trajectory profiles as in Fig. 3b. All in all, there are convincing reasons why rising and falling glides in experiments are not expected to be symmetric, and such asymmetry could explain the fact that the fast rising [i] stands out when verifying Hypothesis 2 above. It must be noted that the 18 fast rising [i] does not similarly stand out using the mixed effects model. Not much can be concluded from the observed numerical values of DR on the two vowel classes. First, the distribution of F1 for [i] glides was not controlled while it is known by the numerical experiment that a relatively small variation of 50 Hz in F affects the mean values of DR. Second, DR is a nonlinear functional, and one has to be careful in interpretations by continuity; e.g., DR may have very large or very small values for large jumps that cross both fA and fB. Third, the jumps never occur instantaneously in the data but rather show an accelerated pitch movement giving DR values closer to 1/3 than in the simulations. Fourth, the dynamics of the pitch following exercise is not understood in such detail that its effect on DR could be evaluated at all. Such dynamics may be a source of systematic error in the estimated DR values, resulting in the lack of verification of the latter part of Hypothesis 2 concerning [A] glides. The computational model predicts modal locking to subharmonics of the vowel formants as well. These are characterized by jumps of one octave, and, if they took place, they would have gone unnoticed in the data processing. Indeed, the pitch detection algorithm admitted only results inside one oc- tave [200 Hz, 400 Hz] which was the nominal gliding range. Otherwise, large jumps were not observed for [A] in the current dataset but such jumps have been observed in other experiments [4]. 7 Conclusions Results from numerical simulations and experiments have been reported on vowel glides. The simulations show a robust and large effect of modal locking of f0 to F1 whenever the frequencies coincide. The experimental data shows a reduced effect of the same kind, supporting the computational model. Hence, experimental results give positive evidence for the existence of significant filter-source feedback in human glide productions even in an experimental setting where the frequency jumps were implicitly asked to be avoided. Moreover, the results suggest that compensatory mechanisms are employed to avoid filter-source feedback induced f0 jumps near the formant frequency F1. Uncontrolled f0 jumps occur frequently in the phonation of boys during the puberty [23]. Changes in the vocal folds and the growing larynx make it difficult for them to produce natural intonation contours or to sing musical melodies. These observations may be partly due to modal locking since the relevant compensatory mechanisms are expected to require re-tuning after a radical change in the geometric dimensions of both the VT and the vocal folds. If the poorly compensated filter-source interaction were a significant cause, then these f0 jumps would have to be more frequent in VT configurations where F1 is low; i.e., in high vowels such as [i] and [u]. 19 Finally, we remark that the active compensation mechanisms to coun- teract the filter-source feedback could perhaps be observed directly. It is expected that especially trained singers would adjust their VT in order to keep the formant and the vocal pitch apart from each other. Acknowledgements The authors were supported by the Finnish graduate school in engineer- ing mechanics, Finnish Academy grant Lastu 135005, the Finnish Academy projects 128204 and 125940, European Union grant Simple4All (grant no. 287678), Aalto Starting Grant, and Abo Akademi Institute of Mathematics. The authors would like to thank O. Engwall for the area function used in the simulations. References [1] Chiba, T. and Kajiyama, M. The Vowel, its Nature and Structure. Tokyo- Kaiseikan Publishing Company Ltd., 1941. [2] Fant, G. The Acoustic Theory of Speech Production. Moulton, The Hague, 1960. [3] Titze, I., Riede, T., and Popolo, P. Nonlinear source-filter coupling in phonation: Vocal exercises, J. Acoust. Soc. Am., 123(4):1902 -- 1915,2008. [4] Zanartu, M., Mehta, D. D., Ho, J. C., Wodicka, G. R., and Hillman, R. E. Observation and anlysis of in vivo vocal fold tissue instabilities produced by nonlinear source-filter coupling: A case study. J. Acoust. Soc. Am. 129:1, 326 -- 339, 2011. [5] Titze, I. Nonlinear source-filter coupling in phonation: Theory, J. Acoust. Soc. Am., 123(5):2733 -- 2749,2008. [6] Hatzikirou H., Fitch W. T., and Herzel H. Voice Instabilities due to Source-Tract Interactions, Acta Acustica united with Acustica, 92:468 -- 475, 2006. [7] Tokuda, I. T., Zemke, M., Kob, M., and Herzel, H. Biomechanical mod- eling of register transitions and the role of vocal tract resonators. J. Acoust. Soc. Am. 127:3, 1528 -- 1536, 2010. [8] Aalto, A., Murtola, T., Malinen, J., Vainio, M. Modal locking between vocal fold and vocal tract oscillations: Simulations in time domain. arXiv:1506.01395, 2015. 20 [9] Aalto, A. A low-order glottis model with nonturbulent flow and mechan- ically coupled acoustic load, Master's thesis, TKK, Helsinki, 2009. Avail- able at http://math.aalto.fi/en/research/sysnum/. [10] Aalto, A., Alku, P., and Malinen, J. A LF-pulse from a Simple Glottal Flow Model, MAVEBA2009 Proceedings, Florence, Italy, 2009. [11] Aalto, A., Aalto, D., Malinen, J., and Vainio, M. 24th Nordic seminar on computational mechanics Proceedings. Helsinki, Finland, 2011. [12] Alku, P. Glottal wave analysis with pitch synchronous iterative adaptive inverse filtering, Speech Communication, 11:109 -- 118,1992. [13] Hor´acek, J., Sidlof, P., and Svec, J. Numerical simulation of self- oscillations of human vocal folds with Hertz model of impact forces, J. Fluids and Structures 20:853 -- 869, 2005. [14] Morse, P. and Ingard, K. Theoretical Acoustics. McGraw Hill, 1968. [15] Hannukainen, A., Lukkari, T., Malinen, J., and Palo, P. Vowel formants from the wave equation, J. Acoust Soc. Am. 122(EL1 -- EL7), 2007. [16] Hor´acek, J. and Svec, J. Aeroelastic model of vocal-fold-shaped vibrat- ing element for studying the phonation threshold, J. Fluids and Struc- tures 16:931 -- 955, 2002. [17] Titze, I. Physiologic and acoustic differences between male and female voices, J. Acoust. Soc. Am., 85(4):1699 -- 1707,1989. [18] Chhetri, D. K., Zhang, Z., and Neubauer, J. Measurement of Young's modulus of vocal folds by indentation, Journal of Voice, 25:1 -- 7, 2011. [19] Pinheiro, J. C. and Bates, D. M. Mixed-effects models in S and S-PLUS, 2000, Springer, New York. [20] Sidlof, P., Hor´acek, J., Ridk´y, V. Parallel CFD simulation of flow in a 3D model of vibrating human vocal folds. Computers & Fluids, 2012, Article in press. [21] Honda, K., Hirai, H, Masaki, S., and Shimada, Y. Role of vertical larynx movement and cervical lordosis in F0 control. Language and Speech, 42:4, 401 -- 411, 1999. [22] Jones, J. A. and Keough, D. Auditory-motor mapping for pitch control in singers and nonsingers. Exp. Brain Res., 190, 279 -- 287, 2008. [23] Boltezar, I. H., Burger, Z. R., and Zargi, M. Instability of voice in adolescence: Pathologic condition or normal developmental variation? J. Periatrics, 130:2, 185 -- 190, 1997. 21
1806.03499
1
1806
2018-06-09T16:10:05
The lock and key model for Molecular Recognition. Is it time for a paradigm shift?
[ "physics.bio-ph", "cond-mat.soft", "q-bio.BM" ]
We review the standard lock and key (LK) model for binding small ligands to larger adsorbent molecule. We discuss three levels of the traditional LK model for binding. Within this model the binding constant or the Gibbs energy of the binding process is related to the total interaction energy between the ligand and the binding site of the adsorbent molecules. When solvent molecules are present, which is the case in all binding processes in biochemistry, we find that a major part of the Gibbs energy of binding could be due to interactions mediated through the solvent molecules. This finding could have major consequences to the applicability of the LK model in drug design, and perhaps require a shift in the prevailing paradigm in this field of research. Keywords: Lock and key model, Binding constant, solvent effect on binding, hydrophilic effect, molecular recognition.
physics.bio-ph
physics
The lock and key model for Molecular Recognition. 1 Is it time for a paradigm shift? Arieh Ben-Naim Department of Physical Chemistry The Hebrew University of Jerusalem Givat Ram, Jerusalem 91904 Israel Abstract We review the standard lock and key (LK) model for binding small ligands to larger adsorbent molecule. We discuss three levels of the traditional LK model for binding. Within this model the binding constant or the Gibbs energy of the binding process is related to the total interaction energy between the ligand and the binding site of the adsorbent molecules. When solvent molecules are present, which is the case in all binding processes in biochemistry, we find that a major part of the Gibbs energy of binding could be due to interactions mediated through the solvent molecules. This finding could have major consequences to the applicability of the LK model in drug design, and perhaps require a shift in the prevailing paradigm in this field of research. Keywords: Lock and key model, Binding constant, solvent effect on binding, hydrophilic effect, molecular recognition. 1. Introduction Binding processes are ubiquitous in biological systems [Ritter (1996), Tropp (1997), Stryer (1975), Ben-Naim (2001, 2010). These range from binding small molecules like oxygen to hemoglobin, various drugs to protein or to DNA, to binding of proteins to DNA [Hard and 2 Lundback (1996), Ptashne (1967), Helene and Lancelot (1982), von Hippel (1994), von Hippel and Goldberger (1979), von Hippel and Berg (1986)]. In all of these the binding mode is highly specific. In this article we discuss the binding of a small ligand to a large macromolecule. This can be extended to the more general process of self-assembly of macromolecules, which we will not discuss here. The idea that binding phenomena are controlled by the co-called Lock and Key (LK) model is quite old. It is attributed to Emil Fischer who postulated this model in 1894. The idea is very simple; the specific action of an enzyme on a substrate can be explained using a Lock and Key analogy. In this analogy, the lock is the enzyme and the key is the substrate. The enzymatic activity can occur only when the correct form of the key (substrate) fits into the key-hole (active site) of the lock (enzyme). The main idea is very simple and easy to understand. A ligand (L) binding to an absorbent molecule (A) will bind more tightly when L "fits" better into the binding site on A. Stillinger and Wasserman (1978), Tabushi and Mizutani (1987), Rebek et al (1987,1988), Lightner et al (1987). Ben-Naim (2001). "Fitting," in its original sense meant geometrical fitting, much as the fitting of a key to a lock, Figures 1. As with real keys and locks, the idea of fitting has gone through a series of variations. In old keys, the shape of the key had to fit the shape of the keyhole for it to be able to work1. Later, the fitting meant to be between a pattern along the key and the counter pattern within the lock. In today's technology, remote control keys do not have any patter fitting at all. The key sends an electronic signal which enables the locking, or the unlocking of the lock. Likewise, the LK model for binding has gone through a series of modifications. From a simple geometrical fitting, Figure 2, to group-pattern fitting, induced fitting and finally, doing away with the fitting altogether. In the next section (2), we shall review the various LK models, and their definitions in terms of the Gibbs energy of binding. In section (3), we shall discuss the new paradigm for binding which does not depend on the geometrical fitting between the ligand and the adsorbent molecule. We shall also discuss some of the consequences of the new paradigm for the theoretical aspects of drug design. 2. The various versions of the LK model 2.1. Purely geometrical fit 3 Figure 3 shows a hard sphere ligand L and a hard adsorbent molecule A having three different sites, 1, 2 and 3. By hard particles we mean that two particles interact via a hard-pair-potential which is infinitely repulsive at distances smaller than some distance 𝜎 and zero everywhere else, Figure 4. In this case, although the ligands "fit" better to site 2, its binding energy is the same for any other site on the surface of A. In other words, the pure geometrical fit does not imply that L will preferentially bind to site 2. The binding Gibbs energy of L to any of the site is zero. 2.2 Geometrical fit with weak interaction energy In most real cases of ligands binding to sites the geometrical fit means maximum interaction energy between the ligand and the groups of the absolvent molecule at the site. In this case the "recognition" of the binding site has been believed to be achieved through direct interaction between the ligand and the site. In this view the selection of the binding site is according to the criterion: 𝑀𝑖𝑛[∆𝑈(𝑖)] (2.1) Where ∆𝑈(𝑖) is the direct interaction between the ligand and the ith site on the polymer A. The minimum is over all possible binding sites, i on A. The simplest means of recognition is through the weak van der Waals interactions. The criterion (2.1) is fulfilled whenever there are more groups on L that interact with groups on A. This is equivalent to the largest area of the surface of contact between L and A, hence the geometrical fit, which characterized the lock-and-key model. Figure 5 shows three ligands L1, L2 and L3, and an adsorbent molecule A having three geometrically different sites, A1, A2 and A3. Without doing any calculation we can correctly guess that ligand L1 will preferentially bind to site A1, ligand L2 will preferentially bind to site A2, and ligand L3 will preferentially bind to site A3. 4 Underlying this guesswork is the assumption that in the absence of the solvent, and for all rigid molecules the Gibbs energy of binding will be reduced to the energy of binding, Ben-Naim (1992, 2001, 2010)i.e. ∆𝐺(𝑏𝑖𝑛𝑑𝑖𝑛𝑔) = ∆𝑈(𝑏𝑖𝑛𝑑𝑖𝑛𝑔) (2.2) Therefore, also the binding constants and the probability of binding to a specific site will be proportional to Pr( 𝑏𝑖𝑛𝑑𝑖𝑛𝑔 𝑡𝑜 𝑎 𝑠𝑝𝑒𝑐𝑖𝑓𝑖𝑐 𝑠𝑖𝑡𝑒) = 𝐶𝑒𝑥𝑝[−𝛽∆𝑈(𝑏𝑖𝑛𝑑𝑖𝑛𝑔)] (2.3) where C is a normalization constant and 𝛽 = (𝑘𝐵𝑇)−1, with 𝑘𝐵 the Boltzmann constant and T the absolute temperature. Clearly, because of equation (2.3) we can identify the geometrical fit with the maximal number of interacting pairs of groups one belonging to the ligand and one belonging to the specific sites, see Figure 6. For instance, ligand L1 would have about 10 group-group interactions with site A1, about 4 interaction energies with site A2 and about 2 interaction energies with site A3. Assuming that each pair interaction contributes the same quantity u to the binding energy, we can conclude that: ∆𝑈(L1on A1 ) = 5𝑢 ∆𝑈(L1on A2 ) = 3𝑢 ∆𝑈(L1on A3 ) = 2𝑢 (2.4) Since u is a negative quantity we can conclude that: Pr(L1on A1) > Pr(L1on A2) > Pr(L1on A3) (2.5) which means that L1 will preferentially bind to A1, less to A2, and less to A3, Figure 6. Similarly, for L2 we can calculate the approximate number of contacts between groups on L2 and groups on the sites of A and conclude that, see Figure 7. Pr(L2on A2) > Pr(L2on A1) > Pr(L2on A3) (2.6) Similarly for L3 we can calculate that it will preferentially bind to A3. 5 This is essentially the physical reason underlying the LK model for binding; geometrical fit means stronger interaction energy, hence also higher probability to bind to a specific site, It should be emphasized that by interaction energies we mean here direct interaction energy between groups on the ligand and groups at the specific site on A. Before we describe the solvent effect of binding we mention two more variations of the LK model which also depend on direct interactions. 2.3 Fitting by means of complimentary functional groups Figure 8 shows a ligand L having three functional groups; a hydroxyl group (OH), and a non-polar group (say methyl) and a positive charged group. The adsorbent molecule has three possible sites each having three functional groups. Clearly, because (+) is attracted to (-), and OH can form a HB with the CO, the ligand will bind preferentially to site A1 rather to either site A2 or A3. The reason is that the charge-charge, and the hydrogen bonding on site A involves the strongest direct interaction energy between L and the site. Hence, we have: Pr(L on A1) > Pr(L on A2) > Pr(L on A3) (2.7) Note however, that geometrical fitting does not play any role here. Only the direct interaction energies between the various groups on the ligand, and on the site determine the interaction energy, hence the preferential probabilities. Another variation of the lock-and-key model is the so-called induced-fit-model, which is essentially the same as the ones described above, but here the fit is achieved after the binding has occurred, Figure 9. The criterion (2.1) may be translated into probabilities, provided the process of binding is carried out in the absence of a solvent. For instance, having two sites (a) and (b), we can say that the probability of binding to the site (a) is larger than the probability of binding to site (b). The relationship between the probability ratio and the difference in the binding energies is (2.8) bUaUbUaUbaexpexpexpPrPr This ratio is also equal to the ratio of the two measurable binding constants. Note that the two carbonyl groups are far apart and therefore the interaction between them is negligible in the 6 absence of a solvent. 3. Molecular recognition through the solvent When the binding occurs in a solvent, the criterion for the selection of the preferable binding site is not (2.1) but instead: 𝑀𝑖𝑛[∆𝐺(𝑖)] (3.1) Where ∆𝐺(𝑖) is the Gibbs energy change for binding to the ith site. The difference between ∆𝐺(𝑖) and ∆𝑈(𝑖) is the solvent-induced contribution to the binding Gibbs energy, and this is related to the solvation Gibbs energies of the three molecules L, A and AL, i.e. 𝛿𝐺(𝑖) = ∆𝐺𝐴𝐿 ∗ (3.2) ∗ − ∆𝐺𝐴 ∗ + ∆𝐺𝐿 In simple non-aqueous solvents the main contributions to 𝛿𝐺(𝑖) comes from the Hard (H) and the Soft (S) parts of the solute-solvent interactions. As we have seen in section 2 these two contributions depend on the volumes and on the surface areas of the three solutes involved in the binding process. It is commonly believed that even when 𝛿𝐺(𝑖) is large compared with ∆𝑈(𝑖), the difference in the values of 𝛿𝐺(𝑖) for different sites i, might not be very large. In other words, the modified criterion (3.1) might be very different from the criterion (2.1), but the difference between the two might not be sensitive to the specific site. Ben-Naim (1980, 1987a, 1987b, 1990,1992, 1994). As an extreme example suppose that for each site i, we have (3.3) Where 𝛿𝐺 is independent of i. In such a case, no matter how large 𝛿𝐺 might be, its effect on the probability ratio for binding to any two sites (a) and (b) would be negligible, i.e. (3.4) GiUiGbUaUbGaGbaexpexpexpexpPrPr 7 As we noted earlier the probability ratio is equal to the ratio of the experimentally measurable binding constants to the sites (a) and (b). Thus, although the binding constant might be considerably modified in the presence of the solvent, the relative preference for the binding site might be the same as in the gaseous phase. This is equivalent to the statement that the lock-and-key model is still valid; the binding Gibbs energy is modified, but the preferential binding site is still determined by the total direct interaction energies ∆𝑈(𝑖). The argument given above, whether made explicitly or implicitly still dominates the thinking of scientists in the pharmaceutical sciences who are engaged in drug design. However, recently it was pointed out that in some binding processes occurring in aqueous solutions, the involvement of hydrophilic (𝐻𝜙𝐼) effects might be so profound that the lock-and-key model might be rendered completely irrelevant to the selection of preferential site. In some cases it can thoroughly modify the way one approaches the problem of drug design, Ben-Naim (1987,2002, 2006, 2009, 2011), Wang and Ben-Naim (1996). We present here one example where solvent-induced effect, based on 𝐻𝜙𝐼 interaction can reverse the preference for the binding sites. Consider again the two sites (a) and (b) in Figure 10. Clearly, the ligand L fits better to site (a) than to site (b). Thus, in accordance with the lock-and-key model site (a) will be preferred over site (b); in probability terms, the ration of the two probabilities is: (3.5) Clearly, in the gaseous phase the ligand will prefer binding to site a. Equivalently, because of the stronger binding energies, the binding constant to site (a) will be larger than the binding constant to site (b). The preferential binding site might be reversed in aqueous solution. Suppose that the ligand L and the polymer A each has a functional group (FG) that can form Hydrogen Bond (HB) with water. Note that these functional groups are not in the binding interface between L and A. Therefore, even in the presence of these FGs, the preferred binding site in the gaseous phase is still 1expexpPrPrbUaUbarg the site (a). This preferred site will probably be maintained if we add an organic solvent, say hexane or benzene. However, in aqueous solutions the whole story is quite different. Suppose that the FGs on L and A are such that when binding to (b), they can be bridged by a water molecule. In this case, the binding Gibbs energy to site (b) will be modified: 8 (3.6) In binding to site (a) the two FGs are far apart so that they do not interact directly or indirectly, therefore the binding Gibbs energy to (a) is: where we have: (3.7) (3.8) The probability ratio in the aqueous solution is: (3.9) For simplicity assuming that 𝛿𝐺(𝑎) is negligible, and that 𝛿𝐺(𝑏) is due to one 𝐻𝜙𝐼 correlation which amount to about −5 𝑘𝐵𝑇 . In this case we shall have (3.10) Thus, although we have started with , i.e. the preferential site is (a), the addition of one 𝐻𝜙𝐼 interaction could change the probability ratio by a factor of 150 in favor of the site (b). Clearly, in such a case, the whole lock-and-key argument becomes irrelevant to the problem of preferential binding site. The finding that 𝐻𝜙𝐼 interactions can change the preferential binding site has far reaching consequences to the problem of drug design, either for designing new drugs or for modifying existing drugs to improve their efficacy. (See for example: Kuntz (1992), Perun and Propst (1989), Propst and Perun (1992), Greer et al (1994) and Roerding and Kroon (1989)) bGbUbGaGaUaGbGaGbGaGrbGaGbargwexpexpexpexpPrPr150gwrr1gr 9 Some specific examples were discussed recently. We shall not present these highly technical examples here. The interested reader should consult the article by Wang and Ben-Naim (1996). In connection with the solvent-induced effect on preferential binding, it should be noted that proteins' surface are very in-homogenous, both in their structure and their chemical constituency. Therefore, it is likely that in most binding processes to protein, both the direct (i.e. lock-and-key model), and the indirect interactions are of comparable magnitudes. There might be crevices on the surface of the proteins that provide tightly fitted binding sites, as well as 𝐻𝜙𝐼interaction through 𝐻𝜙𝐼groups that do not belong to the binding interface. The situation is quite different in DNA, where the surface is much more homogenous, both in terms of the structure as well as in terms of the distribution of FGs. Therefore, in the binding of ligands, drugs or even proteins to DNA, it is more likely that the 𝐻𝜙𝐼interaction will dominate the binding Gibbs energy, and hence also the selection of the binding site. We next turn to a hypothetical example, where we have a seemingly featureless surface, for instance a flat surface with 𝐻𝜙𝐼 groups randomly distributed on it. Looking superficially on this surface, we might not detect any preferred binding site to the ligand L, Figure 11a. Therefore, from the point of view of the lock-and-key model, there exists no preferential binding site. A ligand approaching this surface in water, might "feel" very different affinities to different regions on this apparently homogenous surface. A ligand hovering above the surface might find one region far more favorable for binding than any other region. An illustration of such an example is shown in Figure 11b. Let Pr(0) be the probability of binding of the ligand to an arbitrary region on the surface of the protein which does not contain 𝐻𝜙𝐼groups Examples (1), (2) and (3) in Figure 11. In the absence of a solvente, the Gibbs energy of binding to any point of these regions on the surface of the protein is nearly independent of the location of the binding site. However, in water the ligand might find sites at which the binding Gibbs energy will be much larger due to the formation of one, two or three 𝐻𝜙𝐼interactions, by means of a water-bridge. The relative probabilities for binding to sites (1), (2) and (3), Figure 11 are: 1505exp0Pr1Pr 10 (3.11) It is clear from this example that a seemingly featureless surface might provide binding sites, such as (1), (2) and (3) with significantly different binding constants. We can also imagine that a ligand approaching a DNA molecule might "see" a nearly homogenous surface on the DNA. However, looking through the water, the ligand might "see" some sites which are more preferred for binding than the others. We have discussed here only one type of solvent induced interactions. There are many other possibilities; some of longer range and some of stronger interactions. Longer range can be achieve by chain of two or more molecules figure 12, stronger interactions can be achieved by water molecule bridging more than two functional groups Ben-Naim (2011) 4. Conclusion For over a hundred years the lock and key model for binding a ligand to a site was the only model for binding both small ligands (such as substrate to an enzyme) and large proteins (binding to DNA). Even when a solvent was present, the lock and key model was considered to be valid. The theory of binding did not change. Perhaps the solvent would modify the strength of the interaction between the ligand and the site, hence the binding constant would be affected, but the theory itself was unaffected. Having discovered the importance of the hydrophilic interactions, i.e. water forming hydrogen- bond bridge between two or more hydrophilic groups, the argument based on the lock and key mechanism will not, in some cases be relevant to the binding mechanism. For these cases we suggest that it is time to a paradigm shift from the lock and key model, to specific solvent-induced effect on the binding of a ligand to References Ben-Naim, A. (1980), Hydrophobic Interactions, Plenum Press, New York Ben-Naim, A. (1987a), Solvation Thermodynamics, Plenum Press, New York. 4102.210exp0Pr2Pr6103.315exp0Pr3Pr Ben-Naim, A. (1987b), On the role of water in molecular recognition and self-assembly, Proc. Indian Acad. Sci. 98, (1987) 357. Ben-Naim, A. (1990), Biopolymers, 29, 567 Ben-Naim, A. (1991), Chem. Phys., 93, 8196 11 Ben-Naim, A. (1992), Statistical Thermodynamics for Chemists and Biochemists, Plenum Press, New York Ben-Naim, A.(1994), Hydrophobic–hydrophilic forces in protein folding, in: C.J. Cramer, D.G. Truhlar (Eds.), Structure and Reactivity in Aqueous Solution, ACS, Symposium Series 568, 1994, p.371. Ben-Naim, A. (2001), Cooperativity and Regulation in Biochemical Processes, Kluwer Academic/Plenum Publishers, New York Ben-Naim, A. (2002), Biophys. Chem, 101, 309 Ben-Naim, A. (2006) Molecular Theory of Solutions, Oxford Univ. Press, Oxford University Press, Oxford. Ben-Naim, A. (2009), Molecular Theory of Water and Aqueous Solutions, Part I: Understanding Water, World Scientific, Singapore Ben-Naim, A. (2011), Molecular Theory of Water and Aqueous Solutions, Part II, World Scientific, Singapore Bugg, C.E, Carson, W.M. Montgomery, J.M. Drug design, Scientific American, December (1993). Greer, J., Erickson, J. W., Baldwin, J. J. and Varney, M. D. (1994), Application of the three dimensional structures of protein target molecules in structure-based drugdesign, J. Med. Chem., 37, 1035 Hard, T. and Lundback, T. (1996), Biophysical Chemistry, 62, 121 Helene, C. and Lancelot, G. (1982), Prog Biophys Mol Biol, 39, 1 Kuntz, I. D. (1992), Structure-based strategies for drug design and discovery, Science, 257, 1078 12 Lightner, D. A., Gawronski, J. K. And Wijekoon. W. M. D. (1987), J. Am. Chem. Soc. 109, 923 Perun, T. J., and Propst, C. L. (1989), Computer-Aided Drug Design; Marcel Dekker, Inc., New York Privalov, P. L., Dragan, A. I., Crane-Robinson, C., Breslauner, K. J., Remeta, D. P., and Minetti, C. A. S. A. (2007), J. Mol. Biol. 365, 1 Propst, C. L. and Perun, T. J. (1992), Nucleic Acid Targeted Drug Design; Marcel Dekker, Inc., New York Ptashne, M. (1967), Nature, 214, 232 Rebek, J., Askew, B., Ballester, P. and Costero, A. (1988), J. Am. Chem. Soc. 110, 923 Rebek, J. (1994), Scientific American, 271, 34 Rebek, J. (1988), Molecular recognition: model studies with convergent functional groups, J.Mol.Recognition 1 (1988) 1. Rebek,J. (1987), Model studies in molecular recognition, Science 235 (1987) 1478. Ritter, P. (1996), Biochemistry, A Foundation, Brooks/Cole, Publ. Comp, New York, 1996, p.180. Roerdink, F. H. and Kroon, A. M. (1989), Drug Carrier Systems; John Wiley and Sons, New York Schwabe, J. W. (1997), Curr. Opin. Struct. Biol. 7, 126 Stillinger, F. H. and Wasserman, Z. (1978), J. Phys. Chem. 82, 929 Stryer, L. (1988), Biochemistry, 3rd edition, W. H. Freeman and Company, New York Tabushi, I., and Mizutani, T. (1987), Tetrahedron, 43, 1439 Tropp, B.E. (1997), Biochemistry Concepts and Applications, Brooks/Cole Publ. Comp, New York, 1997, p.179. von Hippel, P. H. (1994), Science, 263, 769 von Hippel, P. H. in R. F. Goldberger ed. (1979), Biological Regulation and Development, Plenum, New York, pp. 279 von Hippel, and Berg, O. G. (1986), Proc Natl Acad Sci USA, 83, 1608 Wang, H. and Ben-Naim, A. (1996), J. Med. Chem., 39, 1531 Wang, H. and Ben-Naim, A. (1997), J. Phys. Chem. B, 101, 1077, 1086 13 14 15 16 17
1201.6325
1
1201
2012-01-30T19:17:02
Origin of Long Lived Coherences in Light-Harvesting Complexes
[ "physics.bio-ph", "physics.chem-ph", "quant-ph" ]
A vibronic exciton model is developed to investigate the origin of long lived coherences in light-harvesting complexes. Using experimentally determined parameters and uncorrelated site energy fluctuations, the model predicts oscillations in the nonlinear spectra of the Fenna-Matthews-Olson (FMO) complex with a dephasing time of 1.3 ps at 77 K. These oscillations correspond to the coherent superposition of vibronic exciton states with dominant contributions from vibrational excitations on the same pigment. Purely electronic coherences are found to decay on a 200 fs timescale.
physics.bio-ph
physics
Origin of Long Lived Coherences in Light-Harvesting Complexes Niklas Christensson1, Harald F. Kauffmann1,2, Tõnu Pullerits3, and Tomáš Mančal4 1Faculty of Physics, University of Vienna, Strudlhofgasse 4, 1090 Vienna, Austria 2Faculty of Physics, Vienna University of Technology, 1040 Vienna, Austria 3Department of Chemical Physics, Lund University, P.O. Box 124, SE-22100 Lund, Sweden and 4Institute of Physics, Faculty of Mathematics and Physics, Charles University, Ke Karlovu 5, Prague 121 16, Czech Republic A vibronic exciton model is developed to investigate the origin of long lived coherences in light- harvesting complexes. Using experimentally determined parameters and uncorrelated site energy fluctuations, the model predicts oscillations in the nonlinear spectra of the Fenna-Matthews-Olson (FMO) complex with a dephasing time of 1.3 ps at 77 K. These oscillations correspond to the coher- ent superposition of vibronic exciton states with dominant contributions from vibrational excitations on the same pigment. Purely electronic coherences are found to decay on a 200 fs timescale. The role of quantum mechanics in biological processes, as well as photosynthetic light-harvesting, has been of interest for a long time [1 -- 3]. The topic received re- newed attention after the observation of long lived oscil- lations in the two-dimensional (2D) spectra of the Fenna- Matthews-Olson (FMO) protein pigment complex [4]. These oscillations were interpreted as a signature of elec- tronic coherences between the delocalized energy eigen- states of the complex, and it was argued that their slow dephasing could enhance the efficiency of energy transfer between the chlorosome antenna and the reaction center [4, 5]. Subsequent studies revealed that the oscillations in FMO have a dephasing time of 1.2 ps at 77 K [6], and that such oscillations are a common feature in light-harvesting complexes [7, 8]. Based on the known structure of FMO (Fig. 1), simulations employing formally exact equations of motions for the reduced density operator found oscilla- tions with dephasing times of 200−300 fs - clearly shorter than the experimental observation [5, 9 -- 11]. Significantly longer dephasing times were found if the transition fre- quency fluctuations (static or dynamic) of the different pigments are assumed to be correlated [12, 13]. However, theoretical studies using molecular dynamics simulations of the interaction between electronic and nuclear degrees of freedom (DOF) have not been able to confirm the ex- istence of such correlations [14], or long lived electronic coherences [15]. To what extent the long lived oscilla- tions in the experiments reflect electronic coherences, or if they influence the transport of energy across the com- plex [4, 16], thus remains an open question. Analysis of excitation dynamics in molecular aggre- gates typically employs a reduced description, where the electronic DOF and their mutual couplings are treated explicitly, and the nuclear modes of the pigments and protein are treated as a heat bath [17, 18]. The pres- ence of underdamped vibrational modes in the bath pro- duces oscillatory signatures in 2D spectra which are simi- lar to the modulations predicted for electronic coherences [19]. The dephasing time of such nuclear coherence are of the order of several picoseconds, and they can often be treated as completely undamped on the time scale of a typical 2D experiment. Low temperature fluores- cence line narrowing (FLN) experiments on FMO have revealed a large number of vibrational modes in the range of 0 − 350 cm−1[20]. However, the oscillations seen in the 2D spectra of FMO cannot be directly related to simple nuclear wavepackets, because the frequency of the oscilla- tions does not match any of the vibrational frequencies, and the Huang-Rhys factors of the modes are too low. On the other hand, recent simulations have revealed un- expected effects on the electronic structure and dynamics if vibrational modes are explicitly included in the system [21]. Motivated by these results, we develop a vibronic exciton Hamiltonian in so called one particle approxima- tion [22, 23] for FMO, in which one vibrational mode on each monomer is treated explicitly. We will show below that this model predicts oscillations in the 2D spectra of FMO with 1.3 ps dephasing times at 77 K, and that these long lived coherences can be traced to superpositions of vibronic exciton states located on the same pigment. The total Hamiltonian of a molecular aggregate in con- tact with the environment is partitioned in a standard way into system, bath and system-bath interaction terms, H = HS +HB +HSB. The system Hamiltonian describes the Qy transition on each bacteriochlorophyll (BChl) in FMO (see Fig. 1) with the vibrational progression of a single vibrational mode. Including the resonance cou- pling between the transitions, the system Hamiltonian reads HS = X n,ν,m,ν ′ [δnmδνν ′ (En+νω0)+Jn,ν;m,ν ′]n, νihm, ν′, (1) where En is the transition frequency of pigment n (site energy), ω0 is the vibrational frequency, and ν is the quantum number of the vibrational mode. The cou- pling energy Jn,ν;m,ν ′ between the individual transitions can be expressed via the electronic resonance coupling Jnm [18] and the Franck-Condon amplitudes of the vi- brational mode [23] Jn,ν;m,ν ′ = hν0ihν′0i. The eigen- values and wave-functions of HS are given by ωα and n,νn, νi, respectively. The bath Hamilto- αi = Pn,ν cα nian, HB, is described as a collection of independent har- monic oscillators, for which the system-bath interaction is given by HSB = Pn,ν ωdn(ω)qnn, νihn, ν. Here q is a generalized coordinate of the environment, and d(ω) is the displacement of the excited state relative to the ground state. We assume that the system-bath Hamilto- nian does not depend on the state of the vibrational mode (ν), implying that a vibrational coherence on an isolated monomer is undamped. Assuming equal but uncorre- lated system-bath interaction for the different pigments, the energy gap correlation function in the local basis is given by Cnm(t) = ω2dndmhqn(t)qm(0)i = C0(t)δnm. When the interaction of the system with the environ- mental modes (i.e. all except those treated explicitly in the system Hamiltonian) is weak, it is advantageous to perform calculations in the eigenstate basis of the system Hamiltonian. The correlation function of the energy gap in the eigenstate representation can be expressed via the expansion coefficients, cα n,ν, and the correlation function of each excitation in the local basis C0(t), n,ν )2 + Pn,ν6=ν ′ n,ν ′)2 + (cα Cαβ(t) = C0(t)nPn,ν(cα n,ν)2(cβ n,ν ′ )2(cβ [(cα n,ν)2(cβ n,ν)2]o = C0(t)γαβ (2) The dephasing dynamics of a coherent superposition be- tween the eigenstates α and β is determined by the line- shape function gαβ(t) = ´ t 0 dτ ′γαβC0(τ ′). The cor- relation function C0(t) is connected to the the spectral density C ′′(ω) via a Fourier transform (see e.g. Ref. [24]). In addition to dephasing, the system-bath inter- action leads to relaxation between the eigenstates of the system. We use Redfield theory [17] (Markov and secular approximation), where the relaxation rate is given by 0 dτ ´ τ kα→β = 2π Pn,ν,m,ν ′ δnmcα n,ν cα m,ν ′ cβ n,ν cβ m,ν ′ ×n(1 + n(ωαβ)) C ′′(ωαβ) + n(−ωαβ) C ′′(−ωαβ)o, (3) where C ′′(ωαβ) > 0 and n(ωαβ) is the Bose-Einstein dis- tribution function. The total relaxation rate from a level α, Γα = 1 2 Pγ6=α kα→γ, determines the lifetime broad- ening of this exciton state. Assuming kBT < ω0, the linear absorption spectrum can be calculated as [18] OD(ω) ∝ ωhX α µα02Re 0 ∞ dt e−gαα(t)−Γαt−i(ωα0−ω)ti∆,Ω, (4) where α runs over all exciton levels and the tran- sition dipole moments µα0 are given by µα0 = n,ν µnhν0i. Here h. . .i∆,Ω denotes the average over Pn,ν cα a random distribution of pigment energies and orienta- tions of complexes. To simulate the oscillations in a third order experiment (i.e. 2D spectra) we adopt the doorway window repre- sentation [25]. Of all Liouville pathways contributing to the signal, only those involving a coherence between two levels in the excited state will give rise to oscillations during the waiting time t2. Without the loss of gener- ality, we focus on the non-rephasing coherence pathways illustrated in Fig. 1, which give rise to oscillations along 2 the diagonal in the non-rephasing 2D spectrum [7]. The response function for this pathway is given by Rαβ,0 = hh(µα0)2(µβ0)2iΩGα(t3)G(2) αβ (t2)Gα(t1)i∆, (5) where Gα(t) = e−iωα0t−Γαt−gαα(t), and G(2) αβ (t2) = e−iωαβ t2 e−gαα(t2)−gββ (t2)+2gαβ (t2)−(Γα+Γβ )t2. In this work we use the site energies (En ) and resonance couplings (Jnm) for FMO Chlorobium tepidum from Ref. [18], and the analytical formula for the overdamped part of the spectral density, C ′′(ω), extracted from a FLN experi- ment [20]. The direction of the transition dipole moments were taken from the protein data bank file 3ENI [26]. FLN experiments have identified 30 vibrational modes in FMO, and the strongest feature in the spectrum arises from three modes around 185 cm−1 [20]. To retain a simple description, we treat this cluster of modes as one effective mode with a frequency of ω0 = 185 cm−1 and a Huang-Rhys factor of 0.05. For calculations with the exciton model, the vibrational mode was included as a underdamped contribution in the spectral density. In all calculations presented in this paper, we sampled the pigment transition energies from a Gaussian distribution with a FWHM of 80 cm−1. The vibronic exciton and exciton model predict very similar linear optical properties as illustrated by the sim- ulated linear absorption spectra shown in Fig. 1. Figure 2(a) shows the time evolution of coherences involving the lowest state of the exciton model. During t2 the signals oscillate with frequencies corresponding to the splitting between the exciton levels and are completely damped af- ter 400 fs. The strong damping can be readily understood from Eqs. (3) and (5). For the exciton model we find that the cross-correlation term gαβ is small and the coherence pathways decay mainly with exp(−gαα(t2) − gββ(t2)). If a (static or dynamic) correlation of the transition fre- quency fluctuations in the site basis is assumed (ad hoc), the cross-correlation functions in the exciton basis be- come larger and enable longer dephasing times. The coherence pathways involving the lowest levels of the vibronic exciton model are shown in Fig. 2(b). The vibronic exciton coherences are remarkably long lived, and the signals show only minor damping on a 2 ps timescale. The long dephasing time of the coherences in the vibronic exciton model can be understood by in- n,ν 2i∆ given in Tab. spection of the expansion factors hcα I and the transition frequency distributions shown in Fig. 1. For instance, state 1 corresponds to 75 % to an excita- tion of the ν = 0 transition of pigment 3, while state 4 has a large contribution of vibrational excitation (ν = 1) on the same pigment. As discussed above, the system-bath interaction is independent of the state of the vibrational mode, and these two vibronic exciton levels will there- fore experience highly correlated fluctuations, resulting in slow dephasing of the 1, 4 coherence. Despite the large contribution from the vibrational excitation to state 4, it has a transition dipole moment which is comparable to that of the other vibronic exciton levels. For non- k3 -k1 α g α β α g g g k2 1 2 6 7 3 5 4 ) t i n u . b r a ( D O 1.0 0.8 0.6 0.4 0.2 0.0 12000 12200 12400 12600 12800 13000 n (cm )-1 Figure 1: Linear absorption spectrum at 77 K for the exci- ton model (black) and the vibronic exciton model (red dash). The gray lines show the distribution of renormalized transi- tion frequencies weighted by the transition strength (scaled by 2/3) for state 1 and 4 in the vibronic exciton model. The filled areas illustrate the relative contribution from electronic (red) and vibronic (blue) excitations on pigment 3. The spec- tra of the vibronic exciton model have been shifted by the reorganization energy of the vibrational mode (−9.25 cm−1) for comparison. The insets show the Feynman diagram il- lustrating the non-rephasing excited state coherence pathway in Eq. (12), and the arrangement of the 7 BChls in FMO (C. tepidum) [26]. This figure was generated using the VMD software [27]. interacting pigments, only the zero-phonon state (ν = 0) has a significant transition dipole moment. The strong transition dipole moment here is the result of intensity borrowing from the electronic transitions on the other pigments. As illustrated in Fig. 1, state 4 of the vi- bronic exciton model exhibits both properties needed for the generation of long lived coherences: a large contribu- tion of vibrational excitation of pigment 3, and a strong transition dipole moment enabled by intensity borrowing. The combination of both effects leads to long lived oscil- lations on the red edge of the linear absorption spectrum as shown in Fig. 2(c). The oscillations show a bi-phasic behavior, where the initial 200 fs decay of the oscillation is due to the decay of coherences between vibronic ex- citon states localized on different pigments (like the 1, 2 coherence, see Fig. 2(b) and Tab. I), while the long lived oscillations reflect coherences between vibronic exciton states localized on the same pigment. A fit to the oscil- lations give a dephasing time of 1.3 ps. By comparing to the stimulated emission signal calculated for the same spectral range, we estimate the modulation of the total signal, including stimulated emission and ground state bleach, to be 5 − 10 % for t2 > 0.3 ps. The Fourier transform of the signal in Fig. 2(c) is shown in Fig. 2(d). The oscillation frequency of 205 cm−1 is higher than ω0 and also higher than the fre- quency observed in the experiment (160 cm−1) [6]. The 3 (b) 0.2 0.1 0.0 -0.1 4 ) l h c B m 1 . 0 ( e d u t i l p m A (a) 0.2 0.1 0.0 -0.1 0.2 0.1 0.0 -0.1 4 ) l h c B m 1 . 0 ( e d u t i l p m A 4 ) l h c B m 1 . 0 ( e d u t i l p m A -0.2 0.0 0.5 1.0 1.5 2.0 t (ps) 2 (c) ) t i n u . b r a ( T F F -0.2 0.0 0.5 1.0 1.5 2.0 t (ps) 2 -0.2 0.0 0.5 1.0 1.5 2.0 t (ps) 2 0.3 (d) 0.2 0.1 0.0 0 100 200 300 400 500 n (cm )-1 Figure 2: Amplitude of the real part of the non-rephasing coherence pathways involving the lowest state. a) RehR1βiΩ,∆ with β = 2 (blue dash), β = 3 (red thin solid), and β = 4 (black solid) for the exciton model. b) RehR1βiΩ,∆ with β = 2 (blue dash), β = 3 (red thin solid), and β = 4 (black solid) for the vibronic exciton model. c) Sum of all non-rephasing coherence pathways, Pαβ RehRαβiΩ,∆ , giving rise to signal in the range 12100 ± 30 cm−1 for the vibronic exciton model with v0 = 185 cm−1 (blue) and v0=117 cm−1 (red) at 77 K. The initial value of the signal is 0.48 and the first minimum at t2 = 0.04 ps has an amplitude of −0.23. The non-rephasing stimulated emission signal calculated for the same parameters and spectral range has an initial value of 0.17. d) Power spectrum of the Fourier transform of the signals in (c) starting from 0.2 ps. oscillation frequency depends on transition energies, elec- tronic couplings and vibrational frequencies according to Eq. (1). Figs. 2(c) and (d) compares the oscillations on the red edge of the spectrum for two different vibrational frequencies found in the FLN experiment, ω0 = 185 cm−1 and ω0 = 117 cm−1. For ω0 = 117 cm−1, the oscillations have a frequency of 140 cm−1 and a shorter dephasing time as compared to ω0 = 185 cm−1. The dephasing time of the coherences depend on the amount of vibra- tional character of the vibronic exciton states, and de- tailed analysis of the oscillations provide information on the energies and nature of the eigenstates not accessible from linear spectra (see Fig. 1). However, the results in Figs. 2 (c) and (d) indicate that more than one vibra- tional mode needs to be included explicitly for a quanti- tative analysis of the oscillations in FMO. In this work we have shown that the vibronic exciton model predicts coherences in FMO with 1.3 ps dephasing times at 77 K. Our model does not invoke static nor dy- namic correlations in the site energies of the pigments, and uses experimentally determined spectral densities and vibrational frequencies. The long lived coherences are found to reflect coherent superpositions of vibronic exciton states with dominant contributions from vibra- hcα n,ν 2i∆ n = 1, ν = 0 n = 1, ν = 1 n = 3, ν = 0 n = 3, ν = 1 n = 4, ν = 0 n = 4, ν = 1 hµ2 BChl) αi∆(µ2 hEαi∆ − hE1i∆(cm−1) α = 1 α = 2 α = 3 α = 4 0.75 0.0 0.0 0.03 0.0 0.2 0.0 0.0 0.51 0.21 0.0 0.0 0.87 0.58 0.0 105 0.58 0.0 0.03 0.2 0.01 0.0 1.3 175 0.15 0.0 0.0 0.67 0.01 0.01 0.57 217 n,ν 2) Table I: Contributions of selected basis excitations (cα to the four first vibronic exciton states averaged over ener- getic disorder. The numbering of the pigments is defined in Fig. 1 and is the same as in Ref. [18]. The two bottom rows show the averaged transition strength in units of the BChl monomer, and the average energy differences between the vibronic exciton levels, respectively. 4 the system as well as on certain bath modes (Eq. (1)), both types of DOF become mixed. This enhances the ef- fective system-bath coupling in a way not accounted for in our exciton model. A similar mixing of system and bath DOF takes place implicitly when the reduced equa- tion of motion for the electronic system is propagated exactly [16]. It is clear that the mixing takes place for all bath modes. However, modes of the protein environ- ment are strongly damped and cannot contribute to a long dephasing time. As shown here, one or more of the underdamped vibrations found in the BChl monomers are needed to account for the enhanced dephasing times of the coherences. Our results imply that the oscillations in the 2D experiment on FMO reflect the dynamics of the nuclear DOF within a single pigment, which should have little impact on the transfer of energy from the chloro- some to the reaction centre. tional excitations on the same pigment. Because vibra- tional modes are an inherent property of all pigments, we expect vibronic excitons to be a general feature in the dy- namics of molecular aggregates. In the exciton language, the long lived coherences reported here correspond to a coherence between the system and the bath. Because the resonance coupling in the vibronic exciton model acts on Acknowledgments: This work was supported by the Wenner-Gren foundation, Österreichischer Austauschdi- enst (WTZ), Austrian Science Foundation (FWF), the Swedish Research Council, KAW foundation, Swedish Energy Agency, the Czech Science Foundation grant GACR 205/10/0989, and the Ministry of Education, Youth, and Sports of the Czech Republic, grant MEB 061107. [1] A. S. Davydov, Biology and Quantum Mechanics (Perg- [14] C. Olbrich, J. Strumpfer, K. Schulten, and U. amon Press, New York, 1981). [2] H. van Amerongen, L. Valkunas, and R. van Gron- delle, Photosynthetic Excitons (World Scientific, Singa- pore, 2000). [3] O. Kühn, V. Sundström, and T. Pullerits, Chem. Phys. 275, 15 (2002). [4] G. S. Engel, T. R. Calhoun, E. L. Read, T.-K. Ahn, T. Mančal, Y.-C. Cheng, R. E. Blankenship, and G. R. Fleming, Nature 446, 782 (2007). [5] A. Ishizaki and G. R. Fleming, Proc. Nat. Acad. Sci. USA 106, 17255 (2009). [6] D. Hayes, J. Wen, G. Panitchayangkoon, R. E. Blanken- ship, and G. S. Engel, Faraday Discuss. 150, 459 (2011). [7] T. R. Calhoun, N. S. Ginsberg, G. S. Schlau-Cohen, Y.- C. Cheng, M. Ballottari, R. Bassi, and G. R. Fleming, J. Phys. Chem. B 113, 16291 (2009). Kleinekathofer, J. Phys. Chem. B 115, 758 (2011). [15] S. Shim, P. Rebentrost, S. Valleau, and A. Aspuru-Guzik, arXiv:1104.2943v2. [16] G. Panitchayangkoon, D. Voronine, D. Abramavicius, J. R. Caram, N. H. C. Lewis, S. Mukamel, and G. S. Engel, Proc. Nat. Acad. Sci. USA 108, 20908 (2011). [17] V. May and O. Kühn, Charge and Energy Transfer Dy- namics in Molecular Systems (Wiley-VCH, Berlin, 2001). [18] J. Adolphs and T. Renger, Biophys. J. 91, 2778 (2006). [19] N. Christensson, F. Milota, J. Hauer, J. Sperling, O. Bixner, A. Nemeth, and H. Kauffmann, J. Phys. Chem. B 115, 5383 (2011). [20] M. Wendling, T. Pullerits, M. A. Przyjalgowski, S. I. E. Vulto, T. J. Aartsma, R. van Grondelle, and H. van Amerongen, J. Phys. Chem. B 104, 5825 (2000). [21] S. Polyutov, O. Kuhn, and T. Pullerits, Chem. Phys. [8] E. Collini, C. Y. Wong, K. E. Wilk, P. M. G. Curmi, P. 394, 21 (2012). Brumer, and G. D. Scholes, Nature 463, 644 (2010). [9] L. Chen, R. Zheng, Y. Jing, and Q. Shi, J. Chem. Phys. 134, 194508 (2011). [22] M. R. Philpott, J. Chem. Phys. 55, 2039 (1971). [23] F. C. Spano, J. Chem. Phys. 116, 5877 (2002). [24] S. Mukamel, Principles of nonlinear optical spectroscopy [10] B. Hein, C. Kreisbeck, T. Kramer, and M. Rodriguez, (Oxford University Press, Oxford, 1995). arXiv:1110.1511v2. [25] W. M. Zhang, T. Meier, V. Chernyak, and S. Mukamel, [11] P. Nalbach, D. Braun, and M. Thorwart, Phys. Rev. E J. Chem. Phys. 108, 7763 (1998). 84, 041926 (2011). [26] D. E. Tronrud, J. Wen, L. Gay, and R. Blankenship, [12] D. Abramavicius and S. Mukamel, J. Chem. Phys. 134, Photosynth. Res. 100, 79 (2009). 174504 (2011). [27] W. Humphrey, A. Dalke, and K. Schulten, J. Mol. Graph- [13] F. Caycedo-Soler, A. W. Chin, J. Almeida, S. F. Huelga, ics 14, 33 (1996). and M. B. Plenio, arXiv:1201.0156v1.
1706.02876
1
1706
2017-06-09T09:46:09
Peripheral neuron survival and outgrowth on graphene
[ "physics.bio-ph" ]
Graphene displays properties which make it appealing for neuroregenerative medicine, yet its interaction with peripheral neurons has been scarcely investigated. Here, we culture on graphene two established models for peripheral neurons: PC12 cells and DRG primary neurons. We perform a nano-resolved analysis of polymeric coatings on graphene and combine optical microscopy and viability assays to assess the material cytocompatibility and influence on differentiation. We find that differentiated PC12 cells display a remarkably increased neurite length on graphene (up to 35%) with respect to controls. DRG primary neurons survive both on bare and coated graphene and present dense axonal networks.
physics.bio-ph
physics
Peripheral neuron survival and outgrowth on graphene Domenica Convertino†,§, Stefano Luin†, Laura Marchetti*,§, and Camilla Coletti*,§ †NEST, Scuola Normale Superiore, Piazza San Silvestro 12, 56127 Pisa, Italy §Center for Nanotechnology Innovation @NEST, Istituto Italiano di Tecnologia, Piazza San Silvestro 12, 56127 Pisa, Italy E-mail: [email protected], [email protected] Keywords: graphene, neuron culture coating, peripheral DRG neuron, PC12, differentiation Abstract Graphene displays properties which make it appealing for neuroregenerative medicine, yet its interaction with peripheral neurons has been scarcely investigated. Here, we culture on graphene two established models for peripheral neurons: PC12 cells and DRG primary neurons. We perform a nano-resolved analysis of polymeric coatings on graphene and combine optical microscopy and viability assays to assess the material cytocompatibility and influence on differentiation. We find that differentiated PC12 cells display a remarkably increased neurite length on graphene (up to 35%) with respect to controls. DRG primary neurons survive both on bare and coated graphene and present dense axonal networks. 1.Introduction A specific feature of peripheral nerves is the ability to spontaneously regenerate after traumatic injuries. In the presence of important gaps where an end-to-end suture is not possible, a surgical approach is used, where nerve conduits (generally, autografts or allografts) are used as bridges between the nerve stumps and provide physical guidance for the axons (1). However, they present limitations in functional recovery and other disadvantages, e.g. size mismatch and increasing healing time for autografts, and rejection and disease transmission for allografts (2). A promising alternative is represented by tissue engineered nerve grafts, that have shown to improve regeneration, reduce scar formation and increase the concentration of neurotrophic factors (1,3). Among materials that can be used for the guide production, silicon stimulates excessive scar tissue formation thus lacking long-term stability, while some other natural polymers, such as collagen and chitosan, lack adequate mechanical and electrical properties (4–6). In recent years, new materials have been suggested as alternative candidates for tissue engineering applications. In particular graphene and other carbon-based nanomaterials have been proposed in life-science applications and nerve tissue regeneration (5,7,8). Graphene is a monolayer of sp2-hybridized carbon atoms arranged in a two-dimensional honeycomb lattice that was first isolated in 2004 from graphite (9). The increasing research interest in graphene is due to its incredible properties: high electron mobility (also at room temperature), superior mechanical properties both in flexibility and strength, high thermal conductivity and high area/volume ratio (10,11). Furthermore, its biocompatibility and chemical stability make it ideally suited for biomedical applications (12). Several studies have used graphene-based materials as biocompatible substrates for growth and differentiation of different cell types, including neural cells (13–16). To date, however, most studies have investigated graphene covalent functionalized forms such as graphene oxide (GO) and its chemical reduction known as reduced graphene oxide (RGO), or liquid phase exfoliated graphene. These graphene-like structures have altered electronic structure and physical properties due to the variable fraction of sp2 and sp3 hybridized carbon atoms. With respect to those graphene-based materials, pristine graphene offers enhanced electrical and tribological properties and most notably an excellent electrical conductivity thus prospecting advantages for nervous system regeneration applications. Indeed, it has been demonstrated that electrical stimulation enhances and directs neurite outgrowth (17,18). To date, the interaction between pristine graphene and peripheral neural cells has been investigated only in two studies (19,20), which suggest a positive effect on neurite outgrowth and proliferation when using graphene coated with fetal bovine serum (FBS). However, in both studies bare glass is used as control, thus the effect on the results of FBS coating, which per se is not a traditional coating for neural cells (21), is not investigated. No detailed study has yet investigated the homogeneity and quality of the coatings typically adopted in neuronal culture. Predicting how polymeric surface coatings distribute onto graphene, due to its hydrophobicity and extreme flatness, is by no means trivial; furthermore, understanding how nerve cells can sense graphene under extracellular-matrix-like coatings is crucially important for possible in vivo applications. Overall, this lack of studies on pristine graphene leaves other carbon-based materials such as carbon nanofibers (CNF), carbon nanotubes (CNT), GO and rGO to star in its play (5,8,15,22). In this work we investigate the potential of graphene as a conductive peripheral neural interface. We select epitaxial graphene obtained via thermal decomposition on silicon carbide (SiC) (23) as the ideal substrate for such investigations. In fact, epitaxial graphene on SiC combines high crystalline quality, scalability, thickness homogeneity and an extreme cleanliness. Graphene is used as a substrate for two cellular models: (i) PC12 cells, a non-neuronal cell line that is able to differentiate upon Nerve Growth Factor (NGF) stimulation and constitutes a widely-used model for peripheral sympathetic neurons (24); (ii) dorsal root ganglion (DRG) sensory neurons, which are used as a model to study regenerative axon growth (25). The homogeneity and quality of a number of polymeric coatings typically adopted for neuronal culturing is investigated, and the most suitable ones are identified and adopted for the reported cultures. Optical microscopy is used to investigate neurite length, number and differentiation while viability assays are used to assess cytocompatibility. We compared results on monolayer graphene on SiC (G) with the ones on 4 possible control substrates: hydrogen etched SiC (SiC), gold coated glass coverslip (Au), glass coverslip (Glass) and polystyrene plate (well). The last two, being routinely used in cell culture procedures, were used as classic controls. SiC controls were implemented since graphene was grown directly on such substrates, which display a good biocompatibility (26) and present prospects for neural implants (27). Finally, gold substrates were used as conductive controls, as they can accelerate axonal elongation applying electrical stimulation (5). 2. Materials and Methods 2.1. Substrates preparation and characterization Graphene on SiC was prepared by adopting a technique which allows to obtain quasi-free standing monolayer graphene (QFMLG) (28). Briefly, buffer layer graphene was obtained via thermal decomposition of on-axis 4H-SiC(0001) performed at 1250 °C in argon atmosphere. QFMLG was obtained by hydrogen intercalating the buffer layer samples at 900 °C in molecular hydrogen at atmospheric pressure (29). The controls adopted in the experiments were: (i) Hydrogen etched SiC(0001) dices – the same substrates were graphene was grown –were cleaned with HF to remove the oxide layer, and hydrogen etched at a temperature of 1250 °C as previously reported (30). (ii) Gold coated glass coverslips were obtained by thermally evaporating on the coverslips, previously cleaned with oxygen plasma, a 2 nm titanium adhesive layer and a 4 nm thin gold layer. (iii) Bare glass coverslips were treated overnight with 65% nitric acid (Sigma). The topography of the samples as well as the graphene number of layers and quality were assessed by both AFM and Raman spectroscopy (Figure S1). 2.2. Surfaces functionalization Samples were coated with different polymeric solutions suggested for the targeted cell cultures and AFM analyses were performed to investigate the morphology of such coatings on graphene and the controls. The following solutions were tested: 0.1 % (w/v) Poly-L-lysine (PLL) solution in H2O (Sigma), 200 µg/ml Collagene Type I (Sigma) in deionized (DI) water, 30 µg/ml Poly-D-lysine (PDL) (Sigma) in PBS, 30 µg/ml PDL and 5 µg/ml laminin (Life Technologies) in PBS. The samples were incubated with the coating solution at 37 °C for 1 hour, 4 hours or 12 hours and rinsed three times in DI water before analyzing their topography via AFM. AFM was performed in tapping mode on samples with and without the polymeric coating, over several areas up to 10x10 μm wide. AFM micrographs were analyzed using the software Gwyddion 2.45. 2.3. PC12 cell culture PC12 cells (ATCC® CRL-1721™) were maintained in a humidified atmosphere at 37 °C, 5% CO2 in RPMI 1640 medium supplemented with 10% horse serum, 5% fetal bovine serum, 1% penicillin/streptomycin and 1% L-glutamine (Gibco). Cells were plated at ∼40-60% confluency onto the substrates previously coated with 0.1 % (w/v) Poly-L-lysine solution (PLL) in H2O (Sigma). Differentiation was achieved using two different procedures: 1) direct addition of 50 ng/ml NGF (Alomone Labs) in complete cell medium after seeding; 2) a 5-6 days priming with 15 ng/ml NGF in complete medium, followed by seeding on the substrates with 50 ng/ml NGF in RPMI medium supplemented with 1% horse serum, 0.5% fetal bovine serum, 1% penicillin/streptomycin and 1% L-glutamine. In both cases, 2/3 of the medium was renewed every 2-3 days. With the second procedure an improved differentiation was observed. The cells were observed at different time points using an inverted microscope equipped with a 20x/40x magnification objective (Leica DMI4000B microscope). Typically, 10 fields per sample were acquired to perform morphometric analysis of PC12 differentiation. Three parameters were measured as previously reported (31): (i) the percentage of differentiated cells (Diff), determined counting the number of cells with at least one neurite with a length equal to or longer than the cell body diameter; (ii) the average number of neurites per cell in the field (av. neurites/cell); (iii) the mean neurite length measuring the longest neurite of each differentiated cell in the field. Cell viability was assessed with the Cell counting Kit-8 assay (CCK-8, Sigma), based on quantification of WST reduction due to the metabolic activity of viable cells. Samples were prepared according to the manufacturer's instructions and measured at the GloMax® Discover multiplate reader (Promega). The results are reported as % over the polystyrene well, considered as control. All the experiments were repeated at least twice independently. 2.4. DRG Cell Culture Rat Embryonic Dorsal Root Ganglion Neurons (R-EDRG-515 AMP, Lonza) cells were maintained in a humidified atmosphere at 37 °C, 5% CO2 in Primary Neuron Basal Medium (PNBM, Lonza) supplemented with L-glutamine, antibiotics and NSF-1 (at a final concentration of 2%) as recommended by the manufacturer. Neurons were plated on the substrates previously coated with a PBS solution of 30 µg/ml Poly-D-lysine (Sigma) (PDL) and 5 µg/ml laminin (Life Technologies). The medium was always supplemented with 100 ng/ml of NGF (Alomone Labs). Since 24h after seeding, 25 µM AraC (Sigma) was added for inhibition of glia proliferation. Half of the medium was replaced every 3-4 days. Neurons were observed at different time points using an inverted microscope (Leica DMI4000B microscope). 3. Results and discussion 3.1. Polymeric coating of epitaxial graphene and control substrates NGF-induced neurite outgrowth of PC12 cells is favored by their adhesion on a substrate. This is typically achieved by coating the dish surfaces with polymers such as poly-L-lysine or biologically derived collagen (24). We applied a water solution of both these coatings to all substrates adopted for our cultures and analyzed by AFM the quality and homogeneity of the coatings after different incubation times, i.e. 1 hour, 4 hours and 12 hours. Panels (a) and (b) of Figure 1 show AFM phase and topography micrographs for the two different coatings and different incubation times on a graphene substrate. Clearly, the Poly-L-Lysine (PLL) coating presents better homogeneity with respect to Collagen Type I coating for which network-like aggregates can be detected. On the other hand, PLL tends to form a homogeneous carpet of spots of 1-2 nm (no aggregates) independent from the incubation time. We also analyzed the same coatings on SiC, gold and glass surfaces. On SiC, PLL and Collagen presented analogous topographies (Figure S2(a) and (b)). Due to the high surface roughness of gold and glass substrates, no conclusions about the quality of the coating could be drawn (Figure S3), although its presence was confirmed by the variation in the hydrophilicity observed with contact angle measurements (Figure S4). Hence, for the PC12 cells cultured in this work, a PLL coating with an incubation time of 4 hours was adopted. The same characterization was performed for the polymeric coatings typically suggested for DRG neurons, i.e., PBS solution of Poly-D-Lysine (PDL) alone and PDL with laminin. Panel (c) in Figure 1 shows the AFM topography and phase images taken for PDL/laminin coated graphene substrates for the three different incubation times (i.e., 1 hour, 4 hours and 12 hours). Also in this case, after the coating, an increased roughness was observed for all time points and in particular the formation of a network-like structure was consistently observed. PDL alone coating gave rise to a similar net (Figure S5(b)). In order to exclude the effect of PBS, we dissolved the same polymeric amount in DI water and after 4h incubation we observed similar structures (Figure S5(a)). On SiC no network formation was observed with or without laminin (Figure S2(c) and (d)). The stability of the coating was confirmed for all the probed incubation times. In this case, PDL with laminin coating (with an incubation time of 4 hours) was selected to carry on the following DRG culture experiments in order to mimic the extracellular matrix. Figure 1. AFM topography images with characteristic line profiles of graphene after three different times of incubation (1 hour, 4 hours and 12 hours) with Collagen Type I coating (200 µg/ml in DI water) (a), 0.1 % (w/v) Poly-L-lysine solution in H2O (b) and Poly-D-Lysine and Laminin coating (30 µg/ml PDL and 5 µg/ml laminin in PBS) (c) (scale bar: 500 nm). The insets show phase images of the same areas. AFM line profile after 4 hours-incubation are shown for each coating. 3.2. Neurite outgrowth of PC12 cell on graphene We first investigated the effect of graphene on PC12 cells. Figure 2(a) reports typical optical micrographs obtained for PC12 cells cultured at day 5 (in the presence and absence of NGF) and at day 7 (with NGF) on the different substrates. The analyses conducted at day 5 evidence that almost no differentiation took place in the absence of NGF, while a significant neurite outgrowth occurred on all substrates upon NGF treatment. Selected morphometric parameters describing the differentiation process were quantified at day 5 and are reported in panels (b), (c) and (d) of Figure 2: the percentage of differentiated cells in the fields (Diff), the average number of neurites per cell (av. neurites/cell) and the length of the longest neurite per differentiated cell (length). This analysis showed that 50% of the cells on graphene differentiate with a mean neurite length of 48.7 µm (panels (b) and (c)). Remarkably, the average length was significantly longer on graphene than on glass (***) and well (**) by 35% and 22% respectively. No significant difference was instead observed in the percentage of differentiation and in the average number of neurites per cell. These results indicate that PC12 cells grow longer neurites on graphene, with a neuronal differentiation that is comparable to that obtained for the standard control wells. Differently from reference (20), we did not observe increased PC12 proliferation on graphene, which could be due to the effect of the FBS coating used in that study. Furthermore, we found that at day 7 living PC12 cells forming neurite networks were present on all the substrates. To better assess graphene cytocompatibility, the viability of undifferentiated PC12 cells was assessed after 3, 5 and 7 days of culture and no statistically significant differences were observed between graphene and the other substrates (Figure 2(e)). Figure 2. (a) PC12 grown on gold (Au), glass coverslip (Glass), graphene (G), SiC and polystyrene (well) coated with PLL in the absence of NGF (first row, scale bar: 50 µm), PC12 cells differentiation at day5 (second row, scale bar: 50 µm) and day 7 (third row, scale bar: 100 µm). Histograms show the quantification of (b) neurite length, (c) percentage of differentiation and (d) average number of neurites per cell after 5 days of NGF treatment of two independent experiments per substrate. For each substrate we analyzed at least 200 cells (nc) from selected fields (nf) (Au: nf=17, nc=203; Glass: nf=20, nc=798; G: nf=20, nc=442; SiC: nf=20, nc=607; well: nf=20, nc=527). (e) Cell viability after 3, 5 and 7 days tested by WST-8. Results are expressed as percentage of cell proliferation relative to the proliferation on the polystyrene control sample. Bars colored as in the other graphs. Data reported as mean ± SE. Nonparametric Kruskal–Wallis test was used for statistical significance, with ** p<0.01, *** p<0.001. 3.3. DRG primary neurons on graphene Next, we investigated the effect of graphene on primary neurons using dorsal root ganglion (DRG) cells while using the same controls adopted in the previous culture. As motivated in 3.1, all the samples were coated with PDL/laminin. Figure 3(a) shows typical optical microscopy images obtained at 1, 4, 9 and 15 days of culture. Starting from day 4, we observed numerous processes and an increase in the cell body area (Figure S6) and in the neurite length (Figure 3(a)). Neurons were observed on all the substrates up to 17 days of culture. We observed that both at day 1 and day 2 the average axon length was higher on graphene than on the other substrates (Figure 3(b)). This observation confirms the trend reported for PC12, although in this case no statistical significance was retrieved. Axonal length was not quantified for longer culturing times due to the highly dense network forming after day 2 (see day 9 and 15 in Figure 3(a)). Given that neuronal growth was previously reported also for non-coated graphene (32,33), we tested also the bare substrates to observe their effect on the neurons. Differently from non-coated glass, where they did not survive, DRG neurons could be nicely cultured on non-coated graphene and gold. On these uncoated substrates, DRG formed cell bodies aggregates and neurite bundles (Figure 3(c)), as previously observed for retinal ganglion cells cultured on graphene (32). Remarkably, DRG neurons survived on uncoated graphene and gold up to 17 days. Concerning material stability issues, it should be noted that graphene showed a good stability and remained intact during the entire culturing period, as revealed by Raman measurements after cell removal (Figure S7). Figure 3. (a) DRG neurons cultured on gold (Au), glass coverslip, graphene (G) and SiC coated with Poly-D- lysine and laminin at different days of culture. Scale bar: 50 µm. (b) Axon length quantification at 24 and 48 hours after cell seeding. We analyzed nf fields for a total of nc cells for each substrate (day1: Au, nf =13, nc=67, Glass: nf =14, nc=75; G: nf =13, nc=29; SiC: nf =12, nc=35; day2: Au, nf=16, nc=89, Glass: nf =13, nc=100, G: nf=12, nc=34, SiC: nf =11, nc=37) and data are reported as mean ± SE. (c) DRG neurons on bare gold and graphene at day 10. Scale bar: 100 µm. 4. Conclusion This work provides novel data about the use of graphene as a substrate for peripheral neuron cultures. We use the PC12 cell line as a consolidated model for peripheral sympathetic neurons and show that such cells grow well on graphene with an increased neurite length (up to 35%) at 5 days of differentiation when compared to controls. Remarkably, graphene performs better than gold, which is an appealing conductive candidate for biomedical applications. Culture of DRG neurons also shows a positive outcome on graphene: neurons survive both on bare and coated graphene until day 17, with a dense axon network that is comparable to the control substrates. In order to investigate graphene influence on axonal outgrowth, further studies are necessary, e.g. using compartmentalized chambers (34). The obtained results confirm the potential of graphene as an active substrate in nerve guidance conduit devices: it would allow the transmission of electrical signals between neurons and make external electrical stimulation feasible to enhance axon regeneration. Acknowledgment The authors would like to thank Fabio Beltram and Giovanni Signore from NEST-Scuola Normale Superiore for fruitful discussion. This project has received funding from the European Union's Horizon 2020 research and innovation programme under grant agreement No. 696656-GrapheneCore1. References 1. Faroni A, Mobasseri SA, Kingham PJ and Reid AJ. Peripheral nerve regeneration: Experimental strategies and future perspectives. Adv Drug Deliv Rev 2015, 82:160–7. 2. Daly W, Yao L, Zeugolis D, Windebank A and Pandit A. A biomaterials approach to peripheral nerve regeneration: bridging the peripheral nerve gap and enhancing functional recovery. J R Soc Interface 2012, 9(67):202–21. 3. Gu X, Ding F and Williams DF. Neural tissue engineering options for peripheral nerve regeneration. Biomaterials 2014, 35(24):6143–56. 4. Tran PA, Zhang L and Webster TJ. Carbon nanofibers and carbon nanotubes in regenerative medicine. Adv Drug Deliv Rev 2009, 61(12):1097–114. 5. Fraczek-Szczypta A. Carbon nanomaterials for nerve tissue stimulation and regeneration. Mater Sci Eng C 2014, 34(1):35–49. 6. Pinho AC, Fonseca AC, Serra AC, Santos JD and Coelho JFJ. Peripheral Nerve Regeneration : Current Status and New Strategies Using Polymeric Materials. Adv Healthc Mater 2016, 5(21):2732– 44. 7. Kostarelos K and Novoselov KS. Exploring the Interface of Graphene and Biology. Science 2014, 344(6181):261–3. 8. Ding X, Liu H and Fan Y. Graphene-Based Materials in Regenerative Medicine. Adv Healthc Mater 2015, 4(10):1451–68. 9. Novoselov KS, Geim AK, Morozov SV, Jiang D, Zhang Y, Dubonos SV, Grigorieva IV and Firsov AA. Electric Field Effect in Atomically Thin Carbon Films. Science 2004, 306(5696):666–9. 10. Castro Neto AH, Guinea F, Peres NMR, Novoselov KS and Geim AK. The electronic properties of graphene. Rev Mod Phys 2009, 81(1):109–62. 11. Lee C, Wei X, Kysar JW and Hone J. Measurement of the of the elastic properties and intrinsic strength of Monolayer Graphene. Science 2008, 321:385–8. 12. Bitounis D, Ali-Boucetta H, Hong BH, Min DH and Kostarelos K. Prospects and challenges of graphene in biomedical applications. Adv Mater 2013, 25(16):2258–68. 13. Agarwal S, Zhou X, Ye F, He Q, Chen GCK, Soo J,Boey F, Zhang H and Peng Chen P. Interfacing live cells with nanocarbon substrates. Langmuir 2010, 26(4):2244–7. 14. Deng M, Yang X, Silke M, Qiu W, Xu M, Borghs G and Chen H. Electrochemical deposition of polypyrrole/graphene oxide composite on microelectrodes towards tuning the electrochemical properties of neural probes. Sensors Actuators, B Chem 2011, 158(1):176–84. 15. Ku SH, Lee M and Park CB. Carbon-Based Nanomaterials for Tissue Engineering. Adv Healthc Mater 2013, 2(2):244–60. 16. Fabbro A, Scaini D, Leon V, Vázquez E, Cellot G, Privitera G, et al. Graphene-Based Interfaces do not Alter Target Nerve Cells. ACS Nano 2015, 10(1):615-23. 17. Schmidt CE, Shastri VR, Vacanti JP and Langer R. Stimulation of neurite outgrowth using an electrically conducting polymer. Proc Natl Acad Sci U S A 1997, 94(17):8948–53. 18. Meng S. Nerve cell differentiation using constant and programmed electrical stimulation through conductive non-functional graphene nanosheets film. Tissue Eng Regen Med 2014, 11(4):274–83. 19. Lee JH, Shin YC, Jin OS, Han D-W, Kang SH, Hong SW and Kim JM. Enhanced neurite outgrowth of PC-12 cells on graphene-monolayer-coated substrates as biomimetic cues. J Korean Phys Soc 2012, 61(10):1696–9. 20. Hong SW, Lee JH, Kang SH, Hwang EY, Hwang YS, Lee MH, Han DW and Park JC. Enhanced neural cell adhesion and neurite outgrowth on graphene-based biomimetic substrates. Biomed Res Int 2014, 2014:1-8. 21. Sun Y, Huang Z, Liu W, Yang K, Sun K, Xing S, et al. Surface coating as a key parameter in engineering neuronal network structures in vitro. Biointerphases 2012, 7(1–4):1–14. 22. Liu X, Miller AL, Park S, Waletzki BE, Zhou Z, Terzic A and Lu L. Functionalized Carbon Nanotube and Graphene Oxide Embedded Electrically Conductive Hydrogel Synergistically Stimulates Nerve Cell Differentiation. ACS Appl Mater Interfaces 2017, 9:14677-90. 23. Starke U, Forti S, Emtsev KV and Coletti C. Engineering the electronic structure of epitaxial graphene by transfer doping and atomic intercalation. MRS Bull 2012, 37(12):1177–86. 24. Greene LA and Tischler AS. PC12 Pheochromocytoma Cultures in Neurobiological Research. Vol. 3. New York: Accademic Press; 1982, 373-414. 25. Chierzi S, Ratto GM, Verma P and Fawcett JW. The ability of axons to regenerate their growth cones depends on axonal type and age, and is regulated by calcium, cAMP and ERK. Eur J Neurosci 2005, 21(8):2051–62. 26. Saddow SE, Frewin CL, Coletti C, Schettini N, Weeber E, Oliveros A and Jarosezski M. Single- crystal silicon carbide: a biocompatible and hemocompatible semiconductor for advanced biomedical applications. Mater Sci Forum 2011, 679–680:824–30. 27. Frewin CL, Locke C, Mariusso L, Weeber EJ and Saddow SE. Silicon carbide neural implants: In vivo neural tissue reaction. Int IEEE/EMBS Conf Neural Eng NER 2013, 661–4. 28. Riedl C, Coletti C, Iwasaki T, Zakharov AA and Starke U. Quasi-free-standing epitaxial graphene on SiC obtained by hydrogen intercalation. Phys Rev Lett 2009, 103(24):1–4. 29. Bianco F, Perenzoni D, Convertino D, De Bonis SL, Spirito D, Perenzoni M, Coletti C, Vitiello MS and Tredicucci A. Terahertz detection by epitaxial-graphene field-effect-transistors on silicon carbide. Appl Phys Lett 2015, 107(13). 30. Frewin CL, Coletti C, Riedl C, Starke U and Saddow SE. A Comprehensive Study of Hydrogen Etching on the Major SiC Polytypes and Crystal Orientations. Mater Sci Forum 2009, 615–617:589– 92. 31. Marchetti L, De Nadai T, Bonsignore F, Calvello M, Signore G, Viegi A, et al. Site-specific labeling of neurotrophins and their receptors via short and versatile peptide tags. PLoS One 2014, 9(11):1–18. 32. Bendali A, Hess LH, Seifert M, Forster V, Stephan AF, Garrido JA and Picaud S. Purified Neurons can Survive on Peptide-Free Graphene Layers. Adv Healthc Mater 2013, 2(7):929–33. 33. Veliev F, Briançon-Marjollet A, Bouchiat V and Delacour C. Impact of crystalline quality on neuronal affinity of pristine graphene. Biomaterials 2016, 86:33–41. 34. Taylor AM, Blurton-Jones M, Rhee SW, Cribbs DH, Cotman CW and Jeon NL. A microfluidic culture platform for CNS axonal injury, regeneration and transport. Nat Methods 2005, 2(8):599–605. Supporting Information Peripheral neuron survival and outgrowth on graphene Domenica Convertino†,§, Stefano Luin†, Laura Marchetti*,§, and Camilla Coletti*,§ †NEST, Scuola Normale Superiore, Piazza San Silvestro 12, 56127 Pisa, Italy §Center for Nanotechnology Innovation @NEST, Istituto Italiano di Tecnologia, Piazza San Silvestro 12, 56127 Pisa, Italy E-mail: [email protected], [email protected] Figure S1. (a) Characteristic AFM topography of an intercalated graphene sample, showing atomically flat terraces separated by steps (scale bar: 400 nm). (b) Raman spectrum of an intercalated graphene sample, obtained using a 532 nm laser and a 50x objective lens. The insert shows the single Lorentzian fitting of the 2D peak, with a narrow FWHM of 28 cm-1. (c) 2D peak position (left) and FWHM (right) distribution in a large area (scale bar: 2 µm). The position and shape of the 2D (~2700 cm-1) peak, originated from a double resonance electron-phonon scattering process, give an indication of the doping and the number of graphene layers. In particular, the single Lorentzian fitting of the peak is characteristic of monolayer graphene, while for bilayer and trilayer graphene the 2D peak becomes broader and the fitting requires multiple Lorentzians. The energy of the peak, blue-shifted with respect to the case of pure undoped graphene, indicates a p- type doping, characteristic of a quasi-free standing monolayer graphene (QFMLG). Figure S2. AFM topography images of SiC samples after three different times of incubation (1 hour, 4 hours or 12 hours) with a coating solution of: (a) PLL, (b) collagen, (c) PDL, (d) PDL/laminin (scale bar: 500 nm). The insets show phase images of the same areas, which are not sensitive to slow changes in height and improve identification of nanometric structures. (e) All the samples are coated with a homogeneous carpet of spots of few nanometers, as showed in the AFM line profile of a SiC sample after 4 hours incubation with PDL/laminin. Figure S3. AFM topography and roughness profiles of gold (a, Au) and nitric acid treated glass (b, Glass) before protein coating and after 4h incubation with Poly-L-lysine (4h PLL) and Collagen Type I (4h COLL) (scale bar: 200 nm). Both the surfaces revealed an initial roughness comparable to the one after the coating, preventing the recognition of nanometric details. Figure S4. Contact angle measurements of gold (Au) and nitric acid treated glass (Glass) before protein coating and after 4h incubation with Poly-L-lysine (PLL) and Collagen Type I (Collagen). All measurements were made using DI water as a probe liquid. Values are the mean ± standard deviation for 3 samples. Non-coated gold was more hydrophobic than non-coated glass. The coatings had opposite effects on the substrates, increasing hydrophilicity for gold and increasing hydrophobicity for glass. Figure S5. (a) AFM topography of graphene samples coated with PDL/laminin dispersed in DI water and PBS after 4h incubation show similar net structures. This implies that the net morphology is independent from the salts in the PBS solution. (b) AFM topography images with a characteristic line profiles of graphene after three different times of incubation (1 hour, 4 hours and 12 hours) with Poly-D-Lysine (PDL) coating (scale bar: 500 nm). Figure S6. Increasing of the cell body area with time in dorsal root ganglion (DRG) cells. For cell soma analysis more than 100 cells per sample were analysed. Cell bodies were approximated to an oval shape and relative areas were evaluated using ImageJ. Figure S7. Raman characterization with 532 nm laser of a graphene sample after cell culture validates the full coverage of graphene. (a) 2D peak position and FWHM distribution in a large area (scale bar: 5 µm). (b) Characteristic Raman spectrum. The maps reveal that the 2D peak and FWHM are very homogeneous across the whole area and the values resemble those measured before the cell culture, with a narrow 2D peak of ~30 cm-1 centered at ~2670 cm-1. Statistical analysis All data are expressed as the average value (mean) ± standard error of the mean (SE) unless stated otherwise. Data were analyzed by using Origin Software and nonparametric Kruskal–Wallis test was used for statistical significance with *p<0.05, ** p<0.01 and *** p<0.001.
1507.00116
1
1507
2015-07-01T06:06:26
Mechanisms of light harvesting by photosystem II in plants
[ "physics.bio-ph" ]
Light harvesting by photosystem II (PSII) in plants is highly efficient and acclimates to rapid changes in the intensity of sunlight. However, the mechanisms of PSII light harvesting have remained experimentally inaccessible. Using a structure-based model of excitation energy flow in 200 nanometer (nm) x 200 nm patches of the grana membrane, where PSII is located, we accurately simulated chlorophyll fluorescence decay data with no free parameters. Excitation movement through the light harvesting antenna is diffusive, but becomes subdiffusive in the presence of charge separation at reaction centers. The influence of membrane morphology on light harvesting efficiency is determined by the excitation diffusion length of 50 nm in the antenna. Our model provides the basis for understanding how nonphotochemical quenching mechanisms affect PSII light harvesting in grana membranes.
physics.bio-ph
physics
Mechanisms of light harvesting by photosystem II in plants Kapil Amarnath,1∗ Doran I. G. Bennett,1∗ Anna R. Schneider,2 Graham R. Fleming1∗ 1Department of Chemistry, University of California, Berkeley, CA and Physical Biosciences Division, Lawrence Berkeley National Lab, Berkeley, CA 2Biophysics Graduate Group, University of California, Berkeley ∗To whom correspondence should be addressed; E-mail: [email protected], [email protected], [email protected] Light harvesting by photosystem II (PSII) in plants is highly efficient and accli- mates to rapid changes in the intensity of sunlight. However, the mechanisms of PSII light harvesting have remained experimentally inaccessible. Using a structure-based model of excitation energy flow in 200 nanometer (nm) x 200 nm patches of the grana membrane, where PSII is located, we accurately simu- lated chlorophyll fluorescence decay data with no free parameters. Excitation movement through the light harvesting antenna is diffusive, but becomes subd- iffusive in the presence of charge separation at reaction centers. The influence of membrane morphology on light harvesting efficiency is determined by the excitation diffusion length of 50 nm in the antenna. Our model provides the basis for understanding how nonphotochemical quenching mechanisms affect PSII light harvesting in grana membranes. 1 Main Text Photosynthetic light harvesting is a multi-scale process that spans from tens of femtoseconds to minutes, and from angstroms to hundreds of nanometers (1). In green plants, photosynthetic light harvesting is performed by photosystems I and II, which are located in separate regions of the thylakoid membrane (2). Photosystem II (PSII), which provides the electrons that drive photosynthesis, is located in the grana stacks of the thylakoid. PSII light harvesting starts when pigments bound to antenna proteins absorb sunlight. The nascent excitation energy is trans- ferred to reaction centers (RC) where it is converted to chemical energy via charge separation. PSII is both an efficient and adaptable light harvesting material. At maximum light harvesting capacity, the average time for excitation capture by PSII (∼300 ps (3, 4)) is significantly less than the excited state lifetime of pigments in the grana (∼2 ns (5, 6)), which results in a >80% efficiency for converting absorbed sunlight into chemical energy (7). In addition, PSII accli- mates to changes in sunlight intensity (8), wavelength (9), and downstream metabolism (10), which occur on the millisecond to minutes timescale. The final outcome of excitation in the grana is determined by the interplay between the rates of excitation energy transfer through the antenna and processes that irreversibly quench excitation (e.g. fluorescence, non-radiative processes, and productive photochemistry in the RC). Time-resolved chlorophyll fluorescence, which is currently the primary technique for measuring PSII light harvesting in a variety of en- vironmental conditions in vivo (11), does not have sufficient spatiotemporal resolution to fully characterize the light harvesting kinetics of PSII (12). Thus a model for PSII light harvesting that accurately captures its underlying dynamics is needed to understand the physical principles that give rise to its functionality. Such principles are essential for understanding the molecular mechanisms of photosynthesis and could be used to engineer artificial light harvesting systems with properties that mimic PSII (13–15). 2 Current models of light harvesting in PSII treat the rates of energy transfer and trapping as parameters whose values are determined by fitting to chlorophyll fluorescence data (4, 16, 17). Conceptually, these models mainly fit within the canonical “lake” and “puddle” models (18), which are currently used to describe the dynamics of energy transfer and trapping by PSII on the hundreds of picoseconds time scale in grana membranes. In the lake model, each excitation tra- verses the grana membrane sufficiently quickly to access all RCs equally. In the puddle model, excitation movement is spatially limited to the smallest photosynthetic unit, which is thought to be a photosystem II supercomplex (PSII-S) and a few surrounding major light harvesting complex II (LHCII) antenna (Fig. 1A, inset) (19). PSII-S consists of two reaction centers and a small number of antenna proteins (20). Up to now, models with free parameters for energy transfer and trapping in PSII have been considered accurate if they can fit chlorophyll fluo- rescence decay data well. However, multiple models that implicitly assume either the lake or puddle model can fit chlorophyll fluorescence decay data equally well (4, 16). As a result, it remains unclear whether the fitted rates found by these methods are physically meaningful and, more broadly, to what extent either the lake or puddle models are appropriate for describing PSII light harvesting. It is well-established that simulating the ultrafast excitation energy trans- fer dynamics in individual pigment-protein complexes requires a structure-based approach that correctly accounts for each chromophore (21, 22). We previously showed that this approach is required even in a small group of pigment-protein complexes in our model of light harvesting within isolated PSII-S (23). Here, we demonstrate that an accurate model for PSII light har- vesting in grana membranes must be based on a correct description of the picosecond dynamics that occur at the nanometer scale of individual pigments. We constructed a parameter-less model for PSII light harvesting that is firmly based on an accurate physical description of the nanoscopic energy transfer dynamics in grana membranes. We performed Monte Carlo simulations on 200 nm x 200 nm patches of the grana membrane 3 (24) (see Methods, below) to generate examples of the mixed (Fig. 1A) and segregated (Fig. 1B) organizations previously observed (2, 26). In the segregated membrane, PSII-S and LHCII separate into so-called PSII-S crystalline arrays and LHCII aggregates. The Monte Carlo model contains a small number of energetic interactions based on in vivo phenomenology and provides plausible locations of LHCII in grana, which are difficult to visualize using existing imaging techniques (24). We superimposed the crystal structures of the chlorophyll pigments in PSII- S (20) and LHCII (27) on these simulations to establish the locations of all of the chlorophylls in a grana patch (Fig. 1A, inset). In previous work on PSII-S, we grouped chlorophylls into highly energetically coupled clusters called domains and demonstrated that it is sufficient to describe the excitation dynamics at the domain level (23). We assume that the kinetics within LHCII and PSII-S are the same as calculated previously on isolated complexes (23). Inhomogeneously averaged rates of energy transfer between domains on different complexes were calculated using Generalized Forster theory (Methods). We modeled photochemistry in the reaction centers with two phenomenological “radical pair” (RP) states. The rate constants between the RP states are the same as those established in previous work on the PSII-S (23). Using our method we can routinely construct a rate matrix for grana membrane patches that contain more pigments (up to 105) than any natural light harvesting system previously modeled. Our model accurately simulates chlorophyll fluorescence data and demonstrates that ex- tracting the amplitude and lifetime components from a simple fit does not capture the complex kinetics of PSII light harvesting. The comparison of the simulation of the mixed membrane (black solid line) with experimental data from thylakoids in ref. (3) (red dotted line) is shown in Fig. 1C. We expected the mixed membrane to be the dominant morphology of the measured thy- lakoids based on the conditions in which the plants (Arabidopsis thaliana) were grown (3, 28). The excellent agreement of the simulation with the experimental data, which was not guaranteed a priori, suggests that our parameter-free model correctly describes excitation energy transport 4 during PSII light harvesting. The simulated decay can be fit well to a sum of a few exponentials (Fig. 1C, inset, green bars), as is frequently done to extract the amplitude and lifetime compo- nents (11). However, the fit does not capture the complex distribution calculated from the rate matrix (Fig. 1C, inset, black bars). Fit-based models that are only sufficiently complex to fit chlorophyll fluorescence decays well will not accurately characterize the underlying PSII light harvesting dynamics. Our simulation of the spatiotemporal dynamics of chlorophyll excitation in the grana al- lowed us to examine the assumptions underpinning the lake and puddle models. Upon uniform initial excitation of either the mixed or segregated membranes, the lake model predicts that the excitation distribution will quickly reach a steady-state spatial distribution. However, we did not observe a steady-state distribution for either the mixed (Movie S1) or segregated (Movie S2) membranes. The puddle model predicts that there are a few (2-3) different types of trapping dynamics at reaction centers throughout the membrane. On the contrary, we observed trapping dynamics (Movies S3 and S4) that vary considerably between different PSII-S. Neither the lake nor the puddle models accurately describe excitation dynamics on the hundreds of picoseconds timescale and hundreds of nanometers length scale of the grana membrane. To understand how excitation moves through the grana, we simulated excitation energy flow from single pigment-protein complexes in LHCII aggregates, PSII crystalline arrays, and mixed membranes (Movies S5-S9). To determine the effect of charge separation on excitation move- ment, we simulated the PSII crystalline arrays and mixed membranes both with and without the radical pair (RP) states in the reaction centers. We quantified excitation transport in these simulations by calculating the time-dependence of the variance of the excitation probability distribution (Fig. 2A). Transport across the grana was well described by fitting the equation σ2(t) − σ2(0) = Atα, (1) 5 Figure 1: Structure-based modeling of energy transfer is required to accurately simu- late PSII light harvesting. (A) and (B) The representative mixed and segregated membrane morphologies generated using Monte Carlo simulations and used throughout this work. PSII supercomplexes (PSII-S) are indicated by the teal discorectangles, while major light harvesting antenna complexes (LHCII) are indicated by the green circles. The segregated membrane forms PSII-S crystalline arrays and LHCII aggregates. As shown schematically in the inset in (A), ex- isting crystal structures of PSII-S (20) and LHCII (27) were overlaid on these membrane patches to establish the locations of all chlorophyll pigments. The teal and green dashed lines outline the excluded area associated with PSII-S and LHCII in the Monte Carlo simulations, respectively. The chlorophyll pigments are indicated in blue, while the protein is depicted by the cartoon ribbon. PSII-S is a 2-fold symmetric dimer of pigment-protein complexes that are outlined by black lines. LHCII-s, CP26, CP29, CP43, and CP47 are antenna proteins, while RC indicates the reaction center. The inhomogeneously-averaged rates of energy transfer between strongly- coupled clusters of pigments were calculated using Generalized Forster theory. (C) Simulated fluorescence decay of the mixed membrane (solid black line) and the PSII-component of exper- imental fluorescence decay data from thylakoid membranes from ref. (3) (red, dotted line). The inset shows the lifetime components and amplitudes of the simulated decay as calculated using our model (black bars) or by fitting to three exponential decays (green bars). 6 where σ2(t) is the variance at time t, σ2(0) is the variance of the initial distribution of excitation, and A and α are fit parameters (Fig. S4B). For α = 1, transport is diffusive, while subdiffusive transport results if α is significantly less than 1. For the LHCII aggregate (Fig. 2A, green tri- angles) and for the mixed membranes and PSII crystalline arrays without RP states (Fig. 2A, red and blue dashed lines, respectively), α ≈ 1 and thus transport within the antenna can be considered diffusive. However, transport for both the mixed and PSII crystal cases with RP states (Fig. 2A, red and blue solid lines, respectively) is sub-diffusive (α < 0.75). Subdiffu- sivity can occur when the energetic differences between sites is on the order of or greater than kBT . We calculated the ∆G for the RC→RP1 (radical pair 1 ) step to be -5.5kBT on the basis of our previously published rates (23). Thus, RP1 serves as an energetic trap that causes sub- diffusive transport. The energy transfer rates in our model are averaged over inhomogeneous realizations, which could mitigate the slow-down of diffusion that occurs when the width of the inhomogeneous distribution is greater than kBT (29). The largest standard deviation of exciton energies across an inhomogeneous distribution in our model is 107 cm−1, which is significantly less than the value of kBT at room temperature (210 cm−1). We note that the diffusive picture of excitation energy transport elaborated here is consistent with predictions made from the PSII-S model (23) and suggests that, in the grana, excitation energy flows neither “directionally” nor energetically downhill on “preferred pathways” (30). We calculated the excitation diffusion length, LD, and the diffusivity, D, to reduce the dynamics observed above to single parameters that can be used to qualitatively understand light harvesting behavior in grana. σ2(t) − σ2(0) = 4Dt = L2 (2) A plot of D as a function of time is shown in Fig. S4A. In the case of subdiffusive transport, the diffusivities decrease in time as excitations are held by the RP1 trap state. The diffusion 7 Figure 2: Excitation transport in grana membranes. Simulation of excitation movement in the five grana membrane configurations shown in the legend: mixed membrane with and without the radical pair (RP) states in the reaction center, PSII-S crystalline array (PSII-S c.a.) with and without the RP states, and LHCII aggregate (LHCII agg). In each case, excitation was initiated on a single pigment-protein complex. (A) The change in the spread of excitation over time. The diffusion exponent α (see text for details) is shown on the right of the plot. (B) Fraction of surviving excitation as a function of net displacement L (eq. 2) from the initial starting point. The dashed line, where the fraction of surviving excitation is 1/e, demarcates the excitation diffusion length (LD). The dimensions of some of the configurations were too small to calculate an LD, so linear extrapolation was used to approximate it (line segments that do not include markers). 8 constants for the three cases ranged from 1−5×10−3 cm2/sec. These values agree with previous experimental work, which used singlet-singlet annihilation measurements of chloroplasts to suggest a lower bound of D ≈ 10−3 cm−2/sec (31). The fraction of excitation remaining as a function of the net spatial displacement, L, is shown in Fig. 2B. LD is defined, by convention, as the minimum net displacement in one dimension achieved by 37% of the excitation population. The LD for the mixed membrane and the PSII crystal with radical pair states were ∼25 nm and ∼15 nm, respectively, though there was some variation depending on the starting point of excitation (Fig. S5). The LD for the LHCII aggregate, mixed membrane, and PSII crystalline array without radical pair states was ∼50 nm. The values of LD and D in the antenna compare favorably with the values observed thus far in organic thin films (32) and quantum dot arrays (29). Using the excitation diffusion length (LD), we explored the longstanding question of how membrane morphology influences the efficiency with which PSII converts absorbed light into chemical energy, the photochemical yield (2, 30). We calculated the photochemical yield in both the mixed and segregated cases to be 0.82 and 0.70, respectively, upon spatially uniform excitation across the membranes. The calculated value of 0.82 for the mixed membrane is in excellent agreement with the estimate of 0.83 derived from chlorophyll fluorescence yield measurements (7). Previous work suggested that the reduced light harvesting efficiency of seg- regated membranes was due to an inability of excitation initiated in LHCII aggregates to drive photochemistry (3), resulting in the “disconnection” of LHCIIs from reaction centers. To ex- plore this hypothesis we initiated excitation on each LHCII in both the mixed and segregated membranes and calculated the resulting photochemical yield, ΦLHCII (Fig. 3A). Both mem- branes have a wide distribution of ΦLHCII (Fig. 3B). (cid:104)ΦLHCII(cid:105) for the segregated membrane, 0.49, is significantly lower than that for the mixed membrane, 0.75. This difference can be explained most simply by the relative diffusion lengths in each membrane environment. In the 9 Figure 3: Influence of grana membrane morphology on photochemical yield. (A) Excitation was initiated at each LHCII in both the mixed (left) and segregated (right) morphologies. The color of the LHCII indicates the fraction of excitation that results in productive photochemistry (ΦLHCII, see colorbar on far right) as simulated with our model. (B) Histograms representing the distribution of ΦLHCII for the mixed and segregated membranes using the coloration from (A). mixed membrane case, nearly all LHCIIs exist within an LD (25 nm) of a PSII-S. In the seg- regated membrane, however, the LHCII aggregate(s) have a diameter comparable to LD (50 nm), resulting in substantial loss of excitation prior to reaching a PSII-S. Clearly, LHCIIs in the segregated membrane show reduced light harvesting function; however, the LHCIIs do not completely disconnect (ΦLHCII ≈ 0). To disconnect an LHCII aggregate from the neighboring PSII-S requires the aggregate diameter to exceed 100 nm, which is unlikely to occur given that a grana membrane disc has a diameter of ∼400 nm (2). The physiological benefit of segre- gation remains unclear and may be explained by other aspects of photosynthesis, as has been proposed (33, 34). Our model accurately describes the energy transfer network of the grana membrane and 10 thus paves the way for understanding PSII light harvesting in all environmental conditions. In intense sunlight, the rate of light absorption by the antenna exceeds the rate of trapping by reaction centers, and the need for photoprotection arises. The mechanisms underlying the non- photochemical quenching of excess excitation in the antenna remain controversial (8). Most of the hypothesized mechanisms of quenching are based, by necessity, on measurements per- formed on isolated pigment-protein complexes (11). However, it has been unclear whether the mechanisms observed in vitro occur in vivo. Our model provides a unified framework for un- derstanding to what extent picosecond dynamics observed in vitro explain in vivo chlorophyll fluorescence data. More broadly, PSII is connected to the rest of photosynthesis through the electron transport chain and the pH gradient across the thylakoid membrane (35,36). Our model could be extended to include the unappressed portion of thylakoid membranes, which contains photosystem I, and integrated into models of the entire thylakoid to fully model photosynthesis in plants. Methods Overview There has been an extensive theoretical development of energy transfer models in photosyn- thetic complexes composed of a few proteins. Excitation dynamics can be calculated directly from the Hamiltonian of a pigment-protein complex (37). However, exact calculations of the dy- namics using the Hamiltonian are computationally costly, and, while feasible on complexes with ∼100 pigments, are impractical for simulating dynamics in the photosystem II-containing por- tions of the thylakoid membrane, which contain hundreds of complexes and >10,000 pigments. Another common technique for calculating excitation dynamics is by perturbative treatments of the Hamiltonian, such as the Generalized Forster and Redfield methods (21). In photosystem II (PSII), the energetic coupling between pigments spans a wide range, such that neither method is 11 appropriate on its own. Modified Redfield theory can interpolate between the strong and weak coupling regimes, but it does not account for dynamic localization in which coupling between the phonon modes of the protein and the pigments’ excited states restrict the delocalization of excitonic states. Therefore, energy transfer in PSII has been treated using a combination of the Modified Redfield and Generalized Forster theories (23,38,39). Transfer within tightly coupled clusters of pigments is treated by Modified Redfield theory and transfer between these clusters is treated using Generalized Forster theory. In previous work on PSII supercomplexes (PSII-S), we used the Modified Redfield/Generalized Forster (MR/GF) method in combination with a novel approach for defining highly coupled clusters of chlorophylls, or domains, that optimizes the separation of timescales between intra- and inter-domain transport (23). Using these domain definitions we were able to coarse-grain the excitation dynamics at the domain level. Recent work using a more exact method (ZOFE) for calculating energy transfer dynamics has further validated our MR/GF simulations (40). Another group has performed HEOM calculations on a quadrant of the PSII-S (41). A compar- ison of our MR/GF simulations with these calculations gives good agreement when appropriate domain definitions are used. We have used the same domain definitions in our simulations of energy transfer in thy- lakoid membranes as were used for modeling PSII-S. We assumed that the timescales of en- ergy transfer between domains on different complexes in the membrane are slow relative to the intra-domain timescales (<1 ps−1) calculated previously (23) and used Generalized Forster theory to calculate the rates between domains on different complexes (see Excitation energy transfer section below). The locations of all of the PSII-S and major light harvesting complex II (LHCII) antenna in a grana membrane were calculated on the basis of Monte Carlo simulations of thylakoid membranes (24) (see Monte Carlo simulations of paired grana membranes section below). The crystal structures of PSII-S (20) and LHCII (27) were overlaid on top of these 12 simulations (see Chlorophyll configurations section below). The dynamics of PSII light harvesting can be represented by a kinetic network composed of 1st-order rate constants. The master equation formalism can be used to calculate the dynamics of excitation: P (t) = KP (t), (3) where K is a rate matrix containing the first order rate constants of excitation transfer between all compartments (which include all domains) in the network, and P (t) is the vector of compart- ment populations (11, 42). Within this framework, we use simple models for charge separation and trapping when excitation reaches the reaction center (see Electron transfer model section below). While there are more complex models for electron transfer within the reaction cen- ter (43), the two-compartment model used here was parameterized on PSII-S with different antenna sizes (23) and is sufficient to determine the overall timescales of this process. The non-radiative rate constant of decay from each domain was (2 ns)−1 (23). The fluorescence rate constant for each exciton was scaled by its transition dipole moment squared with the average fluorescence rate constant across all excitons set to (16 ns)−1 (23). Solving eq. 1 for P (t), which contains the dynamics of excitation population on all com- partments in the network, gives P (t) = CetΛC−1P (0), (4) where C is a matrix which contains the eigenvectors of K, Λ is a diagonal matrix containing the eigenvalues of K, and P (0) is the initial vector of populations. Using eq. 2, we calculate chlorophyll fluorescence decays, the dynamics of excitation over time, and the yields of the different dissipation processes available to excitation - non-radiative decay, fluorescence, and productive photochemistry in the reaction centers (see Calculations of P (t) section below). 13 Monte Carlo simulations of paired grana membranes Grana-scale pigment-protein complex configurations were generated via computer simulations of an extension of the model presented in ref. (24). Briefly, disc-shaped particles L representing LHCII complexes and rod-shaped particles P representing so-called C2S2 PSII-S in 2d layers α and β interacted via hard-core repulsive interactions, plus the attractive energetic potentials ustack(rLα−Lβ ) = −σL σL if rLα−Lβ < σL otherwise, 0 (cid:17)2 −stack (cid:16) rLα−Lβ (cid:40)−M if rLα−Pα < λMσL (cid:40)−agg otherwise, and if rLα−Lα < λaggσL otherwise, 0 0 uM(rLα−Pα) = uagg(rLα−Lα) = (5) (6) (7) with stack = 4 kBT , M = 2 kBT , λagg = 1.15, and all other parameters as in ref. (24). The potentials in Eqs. 5 and 6 are motivated in ref. (24). The square-well attraction in Equation 7 acts a phenomenological, non-specific energetic driving force for LHCII aggregation (44). In the “mixed” condition, agg = 0 kBT , while in the “segregated” condition, agg = 1 kBT . Canonical ensemble Metropolis Monte Carlo simulations of this model were performed as in ref. (24). A ratio of free LHCII to PSII-S particles of 6 was chosen to match the conditions of ref. (3), and a particle packing fraction of 0.75 was chosen to match typical grana conditions. Thus, 64 PSII-S particles and 384 free LHCII particles were initialized in each of two square boxes of side length 200 nm. Simulations were equilibrated with periodic boundary conditions for at least 15M Monte Carlo sweep steps. Chlorophyll configurations Representative configurations from the stroma-side-up layers of the Monte Carlo simulations were selected for analysis. For each configuration, chlorophyll coordinates were assigned for 14 each pigment-protein particle by aligning the center and axis of rotation of the chlorophyll coor- dinates from refs. (20,27) to the center and axis of rotation of the simulated particle (Fig. S3). As discussed and motivated in ref. (23), we have substituted the structure of an LHCII monomer in the place of the minor light harvesting complexes in PSII-S. Because simulated LHCII particles were radially symmetric, an axis of rotation was randomly selected for each LHCII particle. In a small number of cases across the membrane, pigments belonging to one complex enter into the excluded area of a different complex. This overlap originates from a mismatch between the idealized geometries used for the Monte Carlo simulations and the real pigment structure as shown in Fig. S2A. As a result of protein overlap, there exist a small number of analogously high transfer rates within the membrane rate matrix. In order to understand the influence of such rates on the overall description of transport in the membrane, we have removed transfer rates exceeding certain thresholds from the rate matrix and re-calculated the resulting fluorescence decay. As can be seen in Fig. S2B, this does not alter the fluorescence decay dynamics. Excitation Energy Transfer Model The rate of energy transfer from a donor domain d to an acceptor domain a, kdom a←d (eq. 3), is typically calculated for each inhomogeneous realization of a pigment-protein complex using are calculated using the generalized Forster equation (eq. 5-6), where (cid:82) ∞ generalized Forster theory (23, 38, 39). kdom a←d is the Boltzman-weighted (eq. 4) sum of the rates from the excitons in d (M(cid:105) ∈ d) to the excitons in a (N(cid:105) ∈ a). The rates between excitons N (t) is the overlap integral, VM,N2 is the electronic coupling between the two excitons, and Uµ,M is the coefficient of the Mth exciton on the site µ. Calculation of some observable of the 0 dtAM (t)F ∗ system, such as the fluorescence lifetime, is usually done for each realization and then averaged. However, the thylakoid membrane contains >10,000 chlorophylls. Generating hundreds of realizations of the population coefficients and the overlap integrals for the membrane is not 15 only computationally intensive, but may not necessary to accurately describe the dynamics of the system at the protein length scale. kdom a←d = P (d) M = M(cid:105)∈d N(cid:105)∈a kN←M P (d) M (cid:88) (cid:80)M(cid:105)∈d e−EM /kBT (cid:90) ∞ e−EM /kBT kM←N = VM,N2 2π 0 VM,N2 = (cid:88) dtAM (t)F ∗ N (t) Uµ,M Hel µ,γUγ,N2 (8) (9) (10) (11) µ,γ Using the inhomogeneously averaged rate matrix gives the correct overall dynamics and is representative of the energetics of PSII, while greatly increasing the computational efficiency. It is computationally feasible to calculate a single rate matrix for the intact thylakoid mem- brane using MR/GF theory through the use of supercomputers and patience. This calculation, however, is likely to be unnecessary, because a much more computationally efficient method is available for calculating the inhomogeneously averaged rate matrix, which gives the correct overall dynamics, as shown by the fluorescence lifetime comparison in Fig. S1 for a PSII-S. While the dynamics appear to be accurate, this averaging flattens out the energy landscape and will speed up diffusion, as has been shown in quantum dot arrays (29). The standard devi- ation of the exciton energies across the ensemble of inhomogeneous realizations in PSII are significantly less than kBT , suggesting that excitation movement through the grana membrane should not be greatly affected by averaging. Nonetheless, ongoing developments in making exact calculations of energy transfer more scalable (45) will hopefully enable such calculations on membranes in the future. 16 Directly calculating the inhomogeneous average transfer rates between domains substan- tially reduces the computational burden because a single PSII-S contains all of the different domains that occur throughout the grana membrane. The most computationally demanding components for determining the rate matrix for a given inhomogeneous realization are the over- lap integral in eq. 5 and the matrix to transform from the site to the exciton basis in eq. 6. We have assumed that domain definitions, overlap integrals, and the transformation matrices for PSII-S and LHCII are the same at the membrane level as they are in isolated complexes. As a result these terms have been computed previously for several hundred realizations of inho- mogeneous broadening for the four different types of transfers (from LHCII to LHCII, PSII to PSII, LHCII to PSII and PSII to LHCII) in the membrane (23). We tabulated these values and combined them with a calculation of the electrostatic coupling between sites (the only term that must be calculated for each domain-to-domain rate in the membrane). By using our tabulated values for the population matrices and overlap integrals, we have reduced the computational time by 3-4 orders of magnitude, turning an otherwise burdensome calculation into one that can be performed in less than 1 day on a single CPU. We note that this simplification works only because we calculate the inhomogeneously averaged rate between each domain. This allows for a much smaller sampling of the possible combinations of site energies then would be required for even a single inhomogeneous realization of the entire thylakoid membrane. Our algorithm for calculating (cid:104)kdom a←d(cid:105)inhom. real. for each domain pair in the membrane was as follows. We first checked if the two domains were within 60 A of each other. If yes, we proceeded to calculate the average rate; if no, the rate was set equal to 0. We used 60 A as a cutoff based on the distance dependence of rates in the PSII supercomplex. If the two domains were from the same complex, we used the inhomogeneously averaged rate of transfer from already performed generalized Forster/modified Redfield calculations (23). If not, we noted in which type of complex the donor domain and the acceptor domain were in, either LHCII 17 or PSII-S. We calculated kdom a←d using eq. 3-6 with pre-tabulated values of the overlap integral, the population matrices, and the energies of the excitons for a few hundred inhomogeneous realizations. The electronic coupling Hel µ,γ was uniquely calculated for that pair of domains using the ideal dipole approximation. We then averaged over these tabulated inhomogeneous realizations to get (cid:104)kdom a←d(cid:105)inhom. real.. Electron transfer model Both the identity of the primary donor and the kinetics and mechanism of charge separation in the PSII reaction center remain controversial (46). The lack of agreement between the various experimental results and theory has resulted in a variety of phenomenological and conceptual models being used to interpret experimental data (17, 47). In previous work on PSII supercom- plexes, we used the simplest kinetic model that describes the processes known to occur in the PSII reaction center (23), . (12) Here, RC is the reaction center domain composed of the 6 pigments of the reaction center. The “radical pair” states RP1 and RP2 and the rate constants kRC, kCS, and kirr are used to model the electron transfer steps in the reaction center. RP1 and RP2 are non-emissive states that do not have a direct physical analog with charge-separated states in the reaction center. Rather, this approach allows us to describe the overall process of a reversible charge separation step followed by an irreversible step and establish the approximate timescales of these events relative to energy transfer in the light harvesting antenna. The rates were previously parameterized using fluorescence decays of PSII supercomplexes of different sizes (19), with τCS = 0.64 ps, 18 τRC = 160 ps, and τirr = 520 ps (k = 1/τ) (23). Calculations of P (t) Using the definitions established in the previous sections, we can write K as follows: kdom j←i , kcs, krc, kirr, kdump,  Kji = i, j ≤ Ndom, i (cid:54)= j i ∈ RC, j ∈ RP1, j > Ndom i ∈ RP1, j ∈ RC, i > Ndom i ∈ RP1, j ∈ RP2, i, j > Ndom i ≤ Ndom, j = Ncomp − 1 i ≤ Ndom, j = Ncomp kf l i , −(cid:88) k(cid:54)=i kdom k←i, i = j, (13) (14) (15) (16) (17) (18) (19) where Kij is a matrix element of row i and column j, Ndom is the total number of domains, and Ncomp is the total number of compartments (Ndom plus all RP1, RP2, Fl, and dump compart- ments). The sizes of the K for the mixed and segregated membranes modeled (Fig. 1A-B, main text) were both 10882 x 10882, which meant that calculating the eigenvalues and eigenvectors required for simulating the dynamics of excitation (eq. 2) required the use of supercomputers with ∼30 GB of memory. Fluorescence decays were calculated using the equations described in ref. (23). P (0) in all cases was that for ChlA excitation. The photochemical yield was calculated by summing over the populations in all RP2 states at t = 1 sec. References and Notes 1. R. E. Blankenship, Molecular Mechanisms of Photosynthesis (Blackwell Science, 2002). 19 2. J. P. Dekker, E. J. Boekema, Biochimica et Biophysica Acta (BBA)-Bioenergetics 1706, 12 (2005). 3. B. van Oort, et al., Biophysical Journal 98, 922 (2010). 4. A. R. Holzwarth, Y. Miloslavina, M. Nilkens, P. Jahns, Chemical Physics Letters 483, 262 (2009). 5. A. M. Gilmore, T. L. Hazlett, Govindjee, Proc. Natl. Acad. Sci. U. S. A. 92, 2273 (1995). 6. E. Belgio, M. P. Johnson, S. Juri´c, A. V. Ruban, Biophysical journal 102, 2761 (2012). 7. N. R. Baker, Annu. Rev. Plant Biol. 59, 89 (2008). 8. A. V. Ruban, M. P. Johnson, C. D. P. Duffy, Biochimica et Biophysica Acta (BBA) - Bioen- ergetics 1817, 167 (2012). 9. J.-D. Rochaix, Biochimica et Biophysica Acta (BBA)-Bioenergetics 1807, 375 (2011). 10. A. Kanazawa, D. M. Kramer, Proceedings of the National Academy of Sciences 99, 12789 (2002). 11. J. Zaks, K. Amarnath, E. J. Sylak-Glassman, G. R. Fleming, Photosynthesis Research 116, 389 (2013). 12. H. van Amerongen, R. Croce, Photosynthesis Research 116, 251 (2013). 13. R. Croce, H. van Amerongen, Nature Chemical Biology 10, 492 (2014). 14. G. D. Scholes, G. R. Fleming, A. Olaya-Castro, R. van Grondelle, Nature Chemistry 3, 763 (2011). 20 15. G. R. Fleming, G. S. Schlau-Cohen, K. Amarnath, J. Zaks, Faraday Discuss. 155, 27 (2012). 16. K. Broess, G. Trinkunas, A. van Hoek, R. Croce, H. van Amerongen, Biochim. Biophys. Acta, Bioenerg. 1777, 404 (2008). 17. J. Chmeliov, G. Trinkunas, H. van Amerongen, L. Valkunas, Photosynthesis Research pp. in press (available at http://link.springer.com/article/10.1007/s11120–015–0083–3) (2015). 18. G. W. Robinson, Brookhaven Symposia in Biology (1966), vol. 19, p. 16. 19. S. Caffarri, K. Broess, R. Croce, H. vanAmerongen, Biophysical Journal 100, 2094 (2011). 20. S. Caffarri, R. Kouril, S. Kereiche, E. J. Boekema, R. Croce, EMBO J. 28, 3052 (2009). 21. V. I. Novoderezhkin, R. van Grondelle, Physical Chemistry Chemical Physics 12, 7352 (2010). 22. T. Renger, F. Muh, Physical Chemistry Chemical Physics 15, 3348 (2013). 23. D. I. G. Bennett, K. Amarnath, G. R. Fleming, Journal of the American Chemical Society 135, 9164 (2013). 24. A. Schneider, P. Geissler, Biophysical Journal 105, 1161 (2013). 25. See supplementary information for more details. 26. L. A. Staehelin, Photosynthesis Research 76, 185 (2003). 27. Z. Liu, et al., Nature 428, 287 (2004). 28. B. Onoa, et al., PloS one 9, e101470 (2014). 29. G. M. Akselrod, et al., Nano Letters 14, 3556 (2014). 21 30. R. Croce, H. van Amerongen, Journal of Photochemistry and Photobiology B: Biology 104, 142 (2011). 31. C. Swenberg, N. Geacintov, J. Breton, Photochemistry and Photobiology 28, 999 (1978). 32. R. R. Lunt, N. C. Giebink, A. A. Belak, J. B. Benziger, S. R. Forrest, Journal of Applied Physics 105, 053711 (2009). 33. H. Kirchhoff, et al., Biochemistry 46, 11169 (2007). 34. H. Kirchhoff, Philosophical Transactions of the Royal Society B: Biological Sciences 369, 20130225 (2014). 35. X.-G. Zhu, N. R. Baker, D. R. Ort, S. P. Long, et al., Planta 223, 114 (2005). 36. J. Zaks, K. Amarnath, D. M. Kramer, K. K. Niyogi, G. R. Fleming, Proceedings of the National Academy of Sciences 109, 15757 (2012). 37. A. Ishizaki, T. R. Calhoun, G. S. Schlau-Cohen, G. R. Fleming, Physical Chemistry Chem- ical Physics 12, 7319 (2010). 38. V. Novoderezhkin, A. Marin, R. van Grondelle, Physical Chemistry Chemical Physics 13, 17093 (2011). 39. G. Raszewski, T. Renger, J. Am. Chem. Soc. 130, 4431 (2008). 40. J. J. Roden, D. I. Bennett, K. B. Whaley, arXiv preprint arXiv:1501.06674 (2015). 41. C. Kreisbeck, A. Aspuru-Guzik, arXiv preprint arXiv:1502.02657 (2015). 42. C. D. P. Duffy, L. Valkunas, A. V. Ruban, Physical Chemistry Chemical Physics 15, 18752 (2013). 22 43. V. I. Novoderezhkin, E. Romero, J. P. Dekker, R. van Grondelle, ChemPhysChem 12, 681 (2011). 44. M. P. Johnson, et al., The Plant Cell Online 23, 1468 (2011). 45. C. Kreisbeck, T. Kramer, A. Aspuru-Guzik, Journal of Chemical Theory and Computation 10, 4045 (2014). 46. T. Renger, E. Schlodder, ChemPhysChem 11, 1141 (2010). 47. K. Broess, et al., Biophysical Journal 91, 3776 (2006). 48. This research used resources of the National Energy Research Scientific Computing Cen- ter, a DOE Office of Science User Facility supported by the Office of Science of the U.S. Department of Energy under Contract No. DE-AC02-05CH11231. This work was sup- ported by the Director, Office of Science, Office of Basic Energy Sciences, of the U.S. Department of Energy under Contract DE-AC02-05CH11231 and the Division of Chemical Sciences, Geosciences and Biosciences Division, Office of Basic Energy Sciences through Grant DEAC03-76SF000098 (at Lawrence Berkeley National Labs and U.C. Berkeley). 23
1703.04179
1
1703
2017-03-12T21:18:19
The counterbend dynamics of cross-linked filament bundles and flagella
[ "physics.bio-ph" ]
Cross-linked filament bundles, such as in cilia and flagella, are ubiquitous in biology. They are considered in textbooks as simple filaments with larger stiffness. Recent observations of flagellar counterbend, however, show that induction of curvature in one section of a passive flagellum instigates a compensatory counter-curvature elsewhere, exposing the intricate role of the diminutive cross-linking proteins at large-scales. We show that this effect, a material property of the cross-linking mechanics, modifies the bundle dynamics non-trivially, and induces a bimodal $L^2-L^3$ length-dependent material response that departs from the Euler-Bernoulli theory. Hence, the use of simpler theories to analyse experiments can result in paradoxical interpretations. Remarkably, the counterbend dynamics instigates counter-waves in opposition to driven oscillations in distant parts of the bundle, with potential impact on the regulation of flagellar bending waves. These results have a range of physical and biological applications, including the empirical disentanglement of material quantities via counterbend dynamics.
physics.bio-ph
physics
rsos.royalsocietypublishing.org Research Article submitted to journal Subject Areas: xxxxx, xxxxx, xxxx Keywords: xxxx, xxxx, xxxx Author for correspondence: Insert corresponding author name e-mail: [email protected] 7 1 0 2 r a M 2 1 ] h p - o i b . s c i s y h p [ 1 v 9 7 1 4 0 . 3 0 7 1 : v i X r a The counterbend dynamics of cross-linked filament bundles and flagella Rachel Coy1, Hermes Gadêlha2 1CoMPLEX, University College London, London WC1E 6BT, UK. 2Department of Mathematics, University of York, York YO10 SDD, UK. Cross-linked filament bundles, such as in cilia and flagella, are ubiquitous in biology. They are considered in textbooks as simple filaments with larger stiffness. Recent observations of flagellar counterbend, however, show that induction of curvature in one section of a passive flagellum instigates a compensatory counter-curvature elsewhere, exposing the intricate role of the diminutive cross-linking proteins at large-scales. We show that this effect, a material property of the cross-linking mechanics, modifies the bundle dynamics non-trivially, and induces a bimodal L2 − L3 length-dependent material response that departs from the Euler-Bernoulli theory. Hence, theories to analyse experiments can result in paradoxical interpretations. Remarkably, the counterbend dynamics instigates counter-waves in opposition to driven oscillations in distant parts of the bundle, with potential impact on the regulation of flagellar bending waves. These results have a range of physical and biological applications, including the empirical disentanglement of material quantities via counterbend dynamics. the use of simpler 1. Introduction The spontaneous generation of harmonic bending waves along a sperm flagellum has been a source of fascination since it was reported on in the late 17th century [1]. It was not until 1968, however, that the fundamental mechanism behind the flagellar wave propagation was unveiled [2]. ATP-induced inter-microtubule tangential motion is converted into transversal forces that are capable of bending the flagellar assembly altogether, laying the empirical basis for the sliding filament theory for eukaryotic flagellum. c(cid:13) 2014 The Authors. Published by the Royal Society under the terms of the Creative Commons Attribution License http://creativecommons.org/licenses/ by/4.0/, which permits unrestricted use, provided the original author and source are credited. . . . . . . . . . . . . . . . . . . . . . . . 2 r s o s . r o y a s o c e i l t y p u b l i i s h n g . o r g . . . . . . . . . . . . . R . . . . . S o c . . . . . . . o p e n . . s c i . . . . . . . . . . . . . 0 0 0 0 0 0 0 . . Notably, almost one decade before the discovery of the interfilament sliding [2], the existence of such active elements along the sperm flagellum was theorized via a simple fluid- structure interaction model [3]. Machin demonstrated that the combined action of viscous and elastic dissipation experienced by a slender filament rapidly damps any driven oscillation along its length, thus requiring the action of contractile elements in order to sustain large waving amplitude [4]. Later, Goldstein and co-workers [5] elegantly demonstrated that the elastohydrodynamics of any simple Euler-Bernoulli filament moving in a viscous fluid leads to a hyperdiffusive dissipation of bending, characterized by a bending penetration length (cid:96)b, which can be further exploited to extract material parameters from biological filaments in a wide range of length-scales [6]. Hitherto the elastohydrodynamics of active and passive filaments have generated a vast literature of analytical, computational and empirical studies across disciplines [7–16]. Despite the inherent complexity of filament bundles [17–28], as exemplified by the axonemal flagellum [20,21,21], with its 9 + 2 cross-linked microtubule doublets arranged in a cylindrical fashion [20], the textbook elastic bending stiffness has been estimated using a simplistic linear relation between bending moment and curvature [24–28], as derived from Euler-Bernoulli rod theory [29]. Incidentally, the inadequacy of classical rod theories, from Euler-Bernoulli to Timoshenko and Cosserat [29], emerged via paradoxical counterbend empirical responses, Fig. 1(a), first observed by Lindemann and co-workers [30,31], and later captured via a geometrically exact mechanical model by Gadêlha et al [32] (Fig. 1(a)). These studies revealed how the induction of curvature in one section of a passive sperm flagellum instigates non-trivial compensatory counter-curvature elsewhere, namely the counterbend phenomenon [30,31]. They established the critical role of the diminutive elastic linking-proteins while instigating large-amplitude deformations, inherently coupling distant parts along the bundle assembly, despite their small slenderness ratio [30–32]. More recently, the counterbend phenomenon was also exploited in order to extract material quantities from Chlamydomonas flagella [33], despite the relatively short length flagella. The dynamical response of the counterbend phenomenon in passive cross-linked bundles still remains unexplored in the literature. The discovery of counterbend phenomenon highlighted the current need to reassess both the established material measurements [17–19,22–28,34], and the resulting mechanical response, from statics to dynamics, of cross-linked filament bundles [24–28]. The former is crucial in a broad range of biological structures, from the cytoskeleton of eukaryotic cells to cellular division, cross-bridge muscle contraction, and locomotion, via structures like the axoneme. A fundamental challenge, both experimentally and theoretically, is therefore to understand how this complex structure yields bulk material properties and overall cellular mechanical responses and, ultimately, function. In active bundles, the consequences of using inadequate material parameters, which have been used for the past 30 years to investigate flagellar waves [4,8,35–46], are still unknown. This is further confronted with an increasing number of, repeatedly contradicting, active control models for the flagellar wave coordination [4,8,36,37,40,41,43–49]. Paradoxically, in order to induce bending waves, flagellar control models rely on the implementation of filament bundle deformations, in distinct material directions, that are yet to be scrutinized in isolation, from curvature [35,37,42,46] to interfilament sliding [4,8,41,43,49,50], and axial distortions [40,44,47]. This is aggravated by the strong coupling between the unknown activity, and the passive and dissipative components, leading to the non-identifiability of parameters when contrasted against experiments [51,52]. Without the disentanglement between the passive and active elements, and without the rationalization of the resultant mechanical response of cross-linked filament bundles, it is unclear, for example, which competing flagellar control hypothesis [4,8,36,37,40,41,43–48], if any, is able to provide a quantitative understanding of the flagellar regulation and, crucially, function of the internal mechanics and structure. Indeed, any comprehensive model of flagellar bending self-organization depends on reliable measurements of mechanical and material properties of the system in absence of activity [32,33]. . . . . . . . . . . . . . . . . . . . . . . . 3 r s o s . r o y a s o c e i l t y p u b l i i s h n g . o r g . . . . . . . . . . . . . R . . . . . S o c . . . . . . . o p e n . . s c i . . . . . . . . . . . . . 0 0 0 0 0 0 0 . . Figure 1. The counterbend phenomenon and geometry of deformation: (a) Micrograph showing the static configuration of a see urchin sperm rendered passive with its head attached to the coverslip while forced externally by a micro-probe [31], together with the geometrically exact filament-bundle model prediction. Red curve show the model curve fitting result from [32]. (b) 2D representation of the axoneme and the sliding filament mechanism with basal compliance [4,32,34,43]. Micrograph adapted from Gadêlha et al. [32]. Here, we complement the seminal work by Machin [3] and Goldstein [5] on the dynamics of passive filaments, and demonstrate how the nanometric cross-linking proteins that are present in passive cross-linked filament bundles instigate novel dynamical counterbend phenomena. This is in contrast with previous models on flagellar wave coordination [4,8,36,37,40,41,43–49], which incorporate the cross-linking interaction in conjunction with molecular motor dynamics. We consider the dynamical situation in which only the structural passive elements are present. For axonemal filament-bundles, this corresponds to the empirical situation in which molecular motors are rendered passive [30,31,33], Fig. 1(a). The filament-bundle elastohydrodynamical model unveils the occurrence of counter-travelling waves in distant parts of bundle, reducing the propulsive potential of driven oscillations, and even reversing the propulsive direction, from pushing to pulling hydrodynamics. We show that the interplay between the interfilament sliding at the base, and cross-linking dissipation elsewhere, give rise to a bimodal L2 − L3 length-dependent material response that departs from canonical Euler-Bernoulli theory. Hence the use of simpler rod theories to analyse experiments can result in paradoxical interpretations. Furthermore, the counterbend dynamics offers a robust way to measure material quantities empirically, bypassing cumbersome force-displacement experiments at the microscale [27,28,30, 31,33], Fig. 1(a). These results further suggest that the dynamical counter-wave phenomena is likely to play a critical role on the waveform organization, and the subsequent wave direction, of long flagella [45,46,53]. 2. Cross-linked filament bundle elastohydrodynamics We consider a planar representation of cross-linked filament bundles and flagellar axonemes, as depicted in Fig. 1(b), used interchangeably hereafter, composed by two elastic, inextensible filaments that resists deformation with an elastic bending modulus E [4,32,34,43]. Each constituent filament r±(s, t) = r(s, t) ± a/2 n(s, t) is separated by a distance a, much smaller than the filament length a (cid:28) L, normal to the to the centreline r(s, t) at every point in arclength s and time t. Geometry constrains the normal vector n(s, t) = − sin α ex + cos α ey to the plane, where α is the angle between the fixed frame x-axis and the tangent to the centreline t = rs. Like a rail- track [34], the constituent filaments travel distinct contour lengths forcing a geometrical arclength aaatna 4 . . . . . . . . . . . . . r s o s . r o y a s o c e i l t y p u b l i i s h n g . o r g . . . . . . . . . . . . . . . . . . . . . . . R . . . . . S o c . . . . . . . o p e n . . s c i . . . . . . . . . . . . . 0 0 0 0 0 0 0 . . Figure 2. Counter travelling wave formation in filament bundles. (a,b) Euler-Bernoulli hyperdiffusive waveforms and (c,d) filament bundle waveforms for γ = 0, µ = 100. The rescaled sliding displacement ∆/a is overlaid in the waveform in (c,d). The time progression of the waveforms runs from t1 to t5, covering one half of the periodic solution. For the α plots, the time is labelled in multiples of π. mismatch ∆(s, t) = ∆0(t) + a(α(s, t) − α0(t)), where ∆0 and α0 are the length mismatch and tangent angle at s = 0 (Fig. 1)(b). Points of equal contour length along the filament-bundle are connected by elastic links which generate a shearing force, and thus an internal moment, proportional to the sliding displacement f (s, t) = k∆(s, t) with an elastic sliding resistance k. At the basal end, the additional connecting compliance across the filaments, commonly found in 0 f (s(cid:48), t)ds(cid:48) with a spring spermatozoa and inhomogeneous bundles, is Hookean κe∆0(t) = − constant κe [32,43] (Fig. 1(b)). For asymptotically slender filament-bundles, the hydrodynamic forces experienced by an infinitesimal element is anisotropic and linearly related to the local velocity fvis = −ζ⊥(n.rt)n − ζ(cid:107)(t.rt)t, where ζ⊥, ζ(cid:107) are the lowest order resistive coefficients derived from inertialess hydrodynamics [24,25]. Contact forces are not defined constitutively due to the inextensibility constraint [29]. The filament-bundle elastohydrodynamics is governed by the balance of contact forces and contact moments (cid:82) L (2.1) − Eαssss + a2kαss = ζ⊥αt, (cid:82) s L (cid:82) simplified here for small curvatures [5,6,8,43]. The filament-bundle shape is given by the initial 0(cos α(s(cid:48), t), sin α(s(cid:48), t)) ds(cid:48) for an arclength s and time t. value problem r(s, t) = r(0, t) + Boundary conditions ensure the total balance of forces, F (s) = −Eαss + af (s, t), and torques, s f (s(cid:48), t)ds, acting on the bundle [29,32], as detailed in the Supporting M (s) = −Eαs + a Information (SI). The resulting cross-linking mechanics couples distant parts along the bundle 0 (α(s(cid:48)) − via the total momentum balance, now modified non-locally by f (s, t) = −γka α0) ds(cid:48) + ka(α(s) − α0) (SI). Dissipation from different material directions are mediated by the hydrodynamic drag. The filament-bundle dynamics is dictated by the interplay between the elastohydrodynamic hyperdiffusion [5,6] and the cross-linking diffusion [4,34,43]. These boundary moments alter the hyperdiffusion balance in Eq. 2.1 non-trivially, and instigates novel long-range phenomena as we explore below. (cid:82) L (a) The counterbend dynamics: angular actuation The post-transient behaviour of the shape dynamics is captured by single frequency solutions of the form α(s, t) = Re{ α(s)e−iωt}. After convenient rescaling, the eigenvalue problem reduces j=1 Cj erj s−it} which coefficients are to r4 − µr2 − iSp4 = 0, with eigenfunctions α(s, t) = Re{(cid:80)4 . . . . . . . . . . . . . . . . . . . . . . . 5 r s o s . r o y a s o c e i l t y p u b l i i s h n g . o r g . . . . . . . . . Figure 3. The amplitude A of the solutions in the µ = 0 and µ = 100 cases are represented by the grey and black lines respectively; the amplitude of the coloured lines indicates the values of Φ at each point along the flagella as a multiple of π, with the colour denoting the velocity. Note that all velocities below −0.5 and above 0.5 are coloured the same as each of these values respectively. For all of the cases featured in the plot, γ = 0. . . . . R . . . . . S o c . . . . . . . o p e n . . s c i . . . . . . . . . . . . . 0 0 0 0 0 0 0 . . problem reduces to r4 − µr2 − iSp4 = 0, with eigenfunctions α(s, t) = Re{(cid:80)4 determined by non-local boundary moments (SI). After convenient rescaling, the eigenvalue j=1 Cj erj s−it} which coefficients are determined by non-local boundary moments (SI). The dimensionless filament-bundle compliance parameter, also referred as sperm number, Sp = L(ζ⊥ω/E)1/4, captures the battle between elastic and viscous forces [5], whilst the sliding resistance parameter, µ = a2L2k/E, compares bending stiffness with the cross-linking resistance [32,43]. The basal compliance is given by γ = kL/(kL + κe), and varies from γ = 0, corresponding to no interfilament sliding at the base, to γ = 1, for a free basal sliding [32,43]. The emergence of the non-local, counterbend dynamics is depicted in Fig. 2 for a sinusoidal angular actuation of the proximal end (Supplementary Movie 1) with rt(0, t) = 0, and zero force and torque condition at the distal end (SI). An angular amplitude of 0.4362 rad is used to limit the maximum radius of curvature to 0.1. Fig. 2 contrasts Machin's original solutions, Fig. 2(a,b), with the filament-bundle post-transient dynamics, Fig. 2(c,d). Travelling waves originating from the distal end, indicative of the non-local counterbend effect [32], can be clearly seen in Fig. 2(c,d). A relatively small sliding displacement ∆ (overlaid colour in Fig. 2(c,d)), equivalent to only half of the bundle diameter, is capable of deforming the bundle non-locally with the same magnitude of the imposed actuation. The amplitude modulation A(s), phase Φ(s) and velocity of propagation vp(s) = 1/∂sΦ are depicted in Fig. 3, following suitable transformation to solutions of the form α(s, t) = A(s) cos (t − Φ(s)). The abrupt change in wave direction is triggered by the loss in monotonicity of the phase. Non-local counter-waves propagate in opposite direction from the distal end with a non-uniform decaying magnitude along the arclength due to the high-order dissipation. This is in contrast to the one-directional wave propagation of Euler-Bernoulli filaments, µ = 0 in Fig 3. The sharp change in wave direction coincides with reduced wave amplitudes at the point in arclength Fig. 3). Interestingly, the non-local actuation of cross-linking moments at distal parts of the bundle is delayed by the overdamped dynamics, as indicated by the proximal-distal phase difference. Higher Sp causes larger proximal-distal phase mismatch, faster decay of the counterbend wave speed and amplitude, indicative of a destructive interference between proximal and distal waves in Figs. 2(d) and 3 (b), demonstrated by the sharp jump in phase in Fig. 3. . . . . . . . . . . . . . . . . . . . . . . . 6 r s o s . r o y a s o c e i l t y p u b l i i s h n g . o r g . . . . . . . . . Figure 4. Scaling function Υx, representative of the propulsive force. All cases plotted use the sliding resistance parameter value µ = 100, except for the dashed line which plots the case µ = 0. Waveforms corresponding to this µ = 0 case are indicated by ∗. The local maxima/minima Υx(Spm) in the interval Spm ∈ [0, 5] are labelled by black dots. Positions of the values where Υx = 0 are labelled by circles. . . . . R . . . . . S o c . . . . . . . o p e n . . s c i . . . . . . . . . . . . . 0 0 0 0 0 0 0 . . The counterbend dynamics impacts significantly the resulting hydrodynamic propulsion. The time-averaged propulsive force ¯Fx generated over a period may be written as ¯Fx = (ζ⊥ − ζ(cid:107))ω/2π(cid:112)E/ωζ⊥Υx(Sp, µ, γ), where Υx(Sp, µ, γ) is the force scaling function, now modified by the cross-linking dynamics, as depicted in Fig. 4. For an effectively stiff Euler-Bernoulli filament (µ = 0 and low Sp), no propulsive force can be generated [5,54]. This is in accordance with Purcell's scallop theorem where no net propulsion can be achieved via a time reversible motion in inertialess fluids [5,54]. As the basal compliance becomes stiffer, by reducing γ, the cross-linking mechanics switch from a mostly local contribution with small counterbend deformations (γ = 1 in Fig. 4), to a non- local counterbend effect with increasingly large amplitudes at the distal end (γ = 0 in Fig. 4). Ultimately, this causes the propulsive force to vanish (circles in Fig. 4), and even switch the propulsive direction, thus equivalent to a backward net motion. This is despite the imposed waving direction, which is counteracted by waves travelling in opposition at distal parts (Fig. 4). The separatrix in Fig. 5 captures the region in parameter space where the local extrema of Υx(Spm) changes sign. Thus this indicates the region where a significant influence of non-local counterbend effect is predicted. This illustrates how the triad (Sp, µ, γ) may be conveniently tuned to achieve zero, forward or backward propulsion (Fig. 5). Reversal in swimming direction may be achieved by simply increasing the frequency of oscillation for instance. Moreover, the cross-linking dissipation does not affect the bundle penetration length (cid:96)b = L/Sp [5]. Indeed, as Sp increases, only tangential forces contribute to propulsion, thus the positive asymptote for both Euler-Bernoulli filaments and bundles in Fig. 4. The bimodal length-dependent relaxation dynamics The cross-linking mechanics introduces a diffusion-like time-scale, L2ζ⊥/a2k, with a somewhat weaker L2 geometrical dependence, in addition to the high-order, hyperdiffusive scaling L4ζ⊥/E. Indeed, the cross-linking resistance µ = a2L2k/E contrasts the elastohydrodynamic 7 r s o s . r o y a s o c e i l t y p u b l i i s h n g . o r g . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . R . . . . . S o c . . . . . . . o p e n . . s c i . . . . . . . . . . . . . 0 0 0 0 0 0 0 . . Figure 5. Local maxima/minima Υx(Spm), with Spm ∈ [0, 5], across the parameter space (µ, γ). The black line separates the values for which Υx(Spm) is negative from those for which it is positive. bundle, as µ measures the ratio between the natural cross-linking elastic length (cid:96) =(cid:112)E/a2k and cross-linking time-scales. The cross-linking resistance depends on geometrical aspects of the relative to total axial length via µ = (L/(cid:96))2. The cross-linking elastic length (cid:96) has an important biophysical interpretation: it is the dimensional length by which cross-linking effects become prevalent. The cross-linking mechanics become increasingly important when L > (cid:96), while the opposite is found for L ∼ O((cid:96)) or smaller. The latter entails the possibility of studying relaxation countebend phenomena by only varying the length of filament bundle, and motivates rescaling µ relative to the reference length L0, so that µ(L) = L2µ0, µ0 = a2L2 0k/E and L = L/L0, with smallest dimensionless length L = 1. The non-local, counterbend dynamics decaying from initial data, with an amplitude an, is depicted in Fig. 6 for filament-bundles that are clamped at the proximal end (SI). This is characterized by an effective relaxation constant λ−4 nt, with Sn = C1 sin(q1n s) + C2 cos(q1n s) + C3 sinh(q2n s) + C4 cosh(q2n s) for a given participating mode n. The triad of dissipative contributions acts as an effective dispersion medium for both bending and cross-linking deformations, dictated by the same dispersion relation. However, the mode n + (−1)lµ)/2 for shape is captured by the wavenumber-eigenvalue coupling qln = l = 1, 2, which is not only influenced by the effective relaxation constant, reminiscent of pure high-order elastohydrodynamic dissipation, but also by the cross-linking diffusion. Boundary conditions define the transcendental solvability condition for λn, which depend implicitly on both µ and γ, with an infinite number of mode solutions for each parameter set. Curve-fitting expressions for λ1 obtained from numerical solutions are presented in SI. n via α(s, t) =(cid:80) n anSn(s) e−λ4 (cid:113) (cid:112) µ2 + 4λ4 ( The non-local cross-linking diffusion introduces a bimodal length-dependent material response, as illustrated in Fig. 6 for the relaxation time of the fundamental mode (cid:18) L λ1(L) (cid:19)4 . τ1(L) = The filament-bundle relaxation time departs from the characteristic L4 dependence of simpler Euler-Bernoulli filaments. Instead, an L2 asymptote arises for long bundles with γ = 0, while for γ = 1 the transition is to an L3 behaviour. The length-dependent transition between L2 and L3 modes is governed by the basal compliance (Fig. 6 inset (a)). For asymptotically long filament bundles, the exponent of τ1 ∝ Lζ is quadratic, and remains nearly quadratic until γ approaches 1. Such bijection of the material response entails that simultaneous measurement of the bundle mechanical properties, in different material direction, can be extracted from simple relaxation experiments. In particular, increased inter-filament sliding, concentrated towards the clamped end, induces curvature-reversal for long filament bundles (Fig. 6 inset (b)), reminiscent of the 8 . . . . . . . . . . . . . . . . . . . . . . . . . . r s o s . r o y a s o c e l i t y p u b l i i s h n g . o r g . . . . . . . . . . R . . . . . S o c . . . . . . . o p e n s c i . . . . . . . . . . . . . . . . . 0 0 0 0 0 0 0 counterbend phenomenon [32]. Boundary conditions require zero contact forces and torques at the free end, while clamped constraint facilitates the accumulation of cross-linking sliding towards the proximal end. Both bending and cross-linking deformations relax towards the reference configuration with the same effective rate λ1 (Fig. 6 inset (b)). Figure 6. Rescaled relaxation time τ(cid:48) as a function of L(cid:48). The grey lines plotting L2 and L3 are included for comparison. The inset Fig. 5(a) plots the exponent ζ of the relationship τ(cid:48) ∝ Lζ, as calculated for long flagella, against ln γ; Fig. 5(b) shows the progression of the relaxation of the first mode, with the colour indicating the rescaled sliding displacement ∆/a, for the case µ = 100 and γ = 0. Discussion We studied the transient and post-transient dynamics of overdamped filament bundles that are interconnected by linking elastic proteins. Deformations in distinct material directions, arising from the cross-linking interfilament sliding and pure bending deformation, are coupled with local slender-body hydrodynamics. This leads to an effective dispersion mechanism governed by the superposition of short and long-range dissipation mechanisms. Cross-linking stresses are transmitted to distant parts of the bundle via boundary balance of moments. The cumulative moments are able to surpass the high-order elastohydrodynamic dissipation, and shape the bundle structure non-locally, with increased influence for long filament-bundles, or equivalently, large µ. The delicate interplay between the interfilament sliding at the base and the rest of the bundle results in a bimodal dynamic response, which departs from the classical Euler- Bernoulli theory [3,5,29]. When the basal sliding is permitted, cross-linking diffusion is mostly local, and acts to effectively reinforce the bundle structure. Long-range curvature-reversal events, however, are magnified when the basal sliding is constrained [32]. The counterbend dynamics generate spontaneous travelling waves in opposition to driven oscillations, which are capable of suppressing the propulsive potential, and even reverting the direction of propulsion (Fig. 4). Curvature perturbations diffuse more rapidly, a hundred times faster than Euler- Bernoulli hyperdiffusion with an equivalently higher bending rigidity (Fig. 6). Relatively small cross-linking deformations, up to only 30% of the bundle diameter, are capable of exciting large counterbend modes (Fig. 6 inset (b)), and induce a bimodal L2 − L3 length-dependent deviation from the L4-dependence of canonical filaments. Paradoxical measurements may arise 9 r s o s . r o y a s o c e i l t y p u b l i i s h n g . o r g . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . R . . . . . S o c . . . . . . . o p e n . . s c i . . . . . . . 0 0 0 0 0 0 0 . . . . . . . . if standard Euler-Bernoulli theory is used to interpret experiments [26,28,30,31], as exemplified by the paradoxical length-dependent bending stiffness in microtubules [55]. Indeed, the length- deviation predicted here may be mistakenly interpreted as an effective length-dependent bending rigidity via L4ζ⊥/E(L) [5], if rather the Euler-Bernoulli theory is used; de facto, the Euler- Bernoulli theory is traditionally used since the first measurements of flagellar bundles [24,26, 28]. Static, force-displacement experiments that are often used to probe flagellar material quantities [26,28,31,33] are cumbersome, see Fig. 1(a). They require high-precision force calibrated probes and micro-manipulators, and often rely on the rare attachment of the filament's tip to the cover- slip to micro-probe actuation [26,28,31]. This proximity of the filament bundle to the cover-slip can interfere the interfilament sliding due to surface adhesion, biasing in this way force and shape measurements [32]. The counterbend dynamics provides a simpler and robust empirical route for the disentanglement of material parameters. This includes measurements of the basal interfilament elasticity, despite being deeply embedded at the connecting piece of the bundle (Fig. 1). As a result, standard microfluidic designs may be explored to induce shape changes dynamically [15]. Likewise, the dynamical counter-wave phenomenon may also inspire the design of artificial swimmers [56] that are able to reverse the swimming direction by simply increasing, for instance, the frequency of oscillation (see Fig. 4). (cid:112)E/a2k, typically 5µm for flagella [2,4,43,46,57,58]. Interestingly, the majority of eukaryotic The counter-wave phenomena becomes increasingly important for bundles longer than (cid:96) = flagella exceed (cid:96) by few orders of magnitude, from approximately 30µm for Chlamydomonas and sea urchin sperm to almost 200µm for quail sperm [57,58]. Cross-linking effects may also become increasingly important during flagellar growth, and influencing in this way the wave coordination during flagellar reconstitution in Chlamydomonas [59]. Molecular motors organization thus may operate differently for L < (cid:96) and L > (cid:96). Indeed, local flagellar control models [4,46] recently gained empirical support when tested against short flagella experiments [46], a regime where counterbend phenomenon may be negligible (L < (cid:96)). This is despite of the well-known negative support of curvature control models [37,43], tested instead against long flagella (L > (cid:96)). The recurrent contradictions between sliding and curvature control models [4,8,37,43,46,49] may suggest the occurrence of distinct length-dependent flagellar regimes. Linear models coupling the molecular motor reaction kinetics with interfilament sliding spontaneously propagate waves along the axoneme via a Hopf bifurcation [8,38,41,43,45]. The resulting wave train is observed to move from tip to base, i.e. in the direction opposite to what one would expect from a local dissipation theory (when the interfilament sliding is prevented at the base) [8,46,53]. Incidentally, the basal compliance was observed to influence the direction of wave propagation in flagella self-organizations models [43,53], demonstrating the sensitivity of the direction of the travelling wave to details of the connecting piece (basal part) and boundary conditions. This is in agreement with the bimodal response predicted here (see Fig. 2), in which counterbend is maximized when γ = 0 [32]. The wave direction is influenced non-locally by cross- linking effects whose magnitude is regulated by the basal interfilament sliding (see Figs. 2 and 6). This might explain the surprising significance of the basal compliance during the flagellar wave coordination observed recently in empirical studies [57], and even flagellar synchronization that may arise without recurring to hydrodynamical coupling [60]. Nevertheless, the mechanisms by which the flagellar wave direction is selected is poorly understood. Previous studies were reduced to the linear level, and at the nonlinear regime, the dynamical instability generates unstable traveling waves that can propagate in both directions, with potencial for multi-frequency modes [51,61]. The boundary conditions and basal mechanics assist the mode selection nonlinearly, and thus the direction of propagation, emphasizing how the flagellar dynamics is critically dependent on the underlying structural mechanics of the axoneme. Non-local hydrodynamic interactions [58], transversal axonemal deformations [40,44], and geometrical non-linearities [7,35] are also likely to affect the emergence of self-organization in flagellar systems. The high-order diffusive interaction, intrinsic to elastohydrodynamic systems in Eq. 2.1, is observed throughout nature. In non-dilute systems, particles are affected by density variations beyond the nearest neighbours via, for example, biharmonic interactions ut = D1∇2u − D2∇4u [62,63], thus closely related to Eq. 2.1. Despite the relatively short-range influence, such higher- order diffusion instigates non-trivial spatio-temporal dynamics and self-organization in all fields of science [64,65]. They drive instabilities and even mediate the coexistence of spatial patterns and temporal chaos, as observed in Kuramoto-Sivashinsky systems [65]. Other exemplars of local, higher-order diffusion are found in Ginzburg-Landau superconductors, spatial patterning in Cahn-Hilliard and biochemical systems, plus generalized Fisher-Kolmogorov models, water waves and continuum mechanics systems, among others [62,64,65]. In contrast with canonical high-order diffusion systems [62,64,65], the flagellar scaffold, or equivalently, any cross-linked filament bundle immersed in a viscous fluid is governed by high- order dispersion medium that is inherently non-local. Here, the biharmonic diffusion arises instead via a local elastohydrodynamic dissipation [3,5], while the Fickian-like interaction arises through the long-range coupling reflecting the bundle mechanics [4,8,32,34], effectively connecting distant parts of the system via boundary bending moments. This unveils the potential for rich long- range phenomena via reaction-diffusion interactions [8,51,53], from non-local pattern formation to long-range synchronization of auto-oscillators, that are yet to be fully explored in the realm of mathematical biology. We hope that these results will inspire theoreticians and experimentalists to study the dynamical effects of the conterbend phenomenon in filament-bundle as found throughout nature, including prospects for counterbend reaction-diffusion systems in flagellar dynamics, effectively bridging, non-locally, dynamical systems and PDE's. Acknowledgment R.C. thanks Cambridge Bridgwater Summer Research Programme. H.G. acknowledges support by the Hooke Fellowship, University of Oxford, and WYNG Fellowship, Trinity Hall, Cambridge. The authors also thank Dr E.A. Gaffney for enlightening discussions. We dedicate this work in memory of Prof. John R. Blake, whose work and devotion will continue to inspire future generations of scientists. . . . . . . . . . . . . . . . . . . . . . . . 10 r s o s . r o y a s o c e i l t y p u b l i i s h n g . o r g . . . . . . . . . . . . . R . . . . . S o c . . . . . . . o p e n . . s c i . . . . . . . . . . . . . 0 0 0 0 0 0 0 . . Antoine van Leeuwenhoek and the discovery of sperm. Fertility and Sterility, 67(1):16–17, January 1997. STUDIES ON CILIA III. Further Studies on the Cilium Tip and a "Sliding Filament" Model of Ciliary Motility. The Journal of Cell Biology, 39(1):77–94, October 1968. References 1. Stuart S. Howards. 2. Peter Satir. 3. K. E Machin. Wave Propagation Along Flagella. Journal of Experimental Biology, 35(4):796–806, December 1958. 4. Charles J. Brokaw. Flagellar Movement: A Sliding Filament Model. Science, 178(4060):455–462, November 1972. 5. Chris H. Wiggins and Raymond E. Goldstein. Flexive and propulsive dynamics of elastica at low reynolds number. Phys. Rev. Lett., 80:3879, 1998. 6. C. H. Wiggins, D. Riveline, A. Ott, and R. E. Goldstein. Trapping and wiggling: elastohydrodynamics of driven microfilaments. Biophys J, 74(2 Pt 1):1043–1060, Feb 1998. 7. H. Gadêlha, E. A. Gaffney, D. J. Smith, and J. C. Kirkman-Brown. 11 . . . . . . . . . . . . . . . . . . . . . . . r s o s . r o y a s o c e i l t y p u b l i i s h n g . o r g . . . . . . . . . . . . . R . . . . . S o c . . . . . . . o p e n . . s c i . . . . . . . . . . . . . 0 0 0 0 0 0 0 . . Nonlinear instability in flagellar dynamics: a novel modulation mechanism in sperm migration? Journal of The Royal Society Interface, 7:1689–, 2010. 8. SÃl'bastien Camalet, Frank Jülicher, and Jacques Prost. Self-organized beating and swimming of internally driven filaments. Phys. Rev. Lett., 82:1590, 1999. 9. L. Bourdieu, T. Duke, M. B. Elowitz, D. A. Winkelmann, S. Leibler, and A. Libchaber. Spiral defects in motility assays: A measure of motor protein force. Phys. Rev. Lett., 75:176–179, 1995. 10. R. E. Goldstein and S. A. Langer. Nonlinear dynamics of stiff polymers. Phys. Rev. Lett., 75:1094–1097, 1995. 11. H. C. Fu, C. W. Wolgemuth, and T. R. Powers. Beating patterns of filaments in viscoelastic fluids. Phys. Rev. E, 78:041913–041925, 2008. 12. T. S. Yu, , E. Lauga, and A. E. Hosoi. Experimental investigations of elastic tail propulsion at low reynolds number. Phys. Fluids, 18:0917011–0917014, 2006. 13. Sarah D. Olson, Sookkyung Lim, and Ricardo Cortez. Modeling the dynamics of an elastic rod with intrinsic curvature and twist using a regularized Stokes formulation. Journal of Computational Physics, 238:169–187, April 2013. 14. A. K. Tornberg and M. J. Shelley. Simulating the dynamics and interactions of flexible fibers in stokes flows. J. Comput. Phys., 196:8–40, 2004. 15. Vasily Kantsler and Raymond E. Goldstein. Fluctuations, Dynamics, and the Stretch-Coil Transition of Single Actin Filaments in Extensional Flows. Physical Review Letters, 108(3):038103, January 2012. 16. Tim Sanchez, Daniel T. N. Chen, Stephen J. DeCamp, Michael Heymann, and Zvonimir Dogic. Spontaneous motion in hierarchically assembled active matter. Nature, 491(7424):431–434, November 2012. 17. Claus Heussinger, Felix Schüller, and Erwin Frey. Statics and dynamics of the wormlike bundle model. Phys. Rev. E, 81(2):021904, Feb 2010. 18. M. M. A. E. Claessens, C. Semmrich, L. Ramos, and A. R. Bausch. Helical twist controls the thickness of f-actin bundles. Proceedings of the National Academy of Sciences, 105(26):8819–8822, 2008. 19. Mireille MAE Claessens, Mark Bathe, Erwin Frey, and Andreas R Bausch. Actin-binding proteins sensitively mediate f-actin bundle stiffness. Nature materials, 5(9):748–753, 2006. 20. Don W. Fawcett, William Bloom, and Elio Raviola. A Textbook of Histology. Chapman & Hall, June 1994. 21. B. Afzelius. Electron microscopy of the sperm tail. Biophys. Cytol., 5:269, 1959. 22. JA Tolomeo and MC Holley. Mechanics of microtubule bundles in pillar cells from the inner ear. Biophysical journal, 73(4):2241, 1997. 23. Itsushi Minoura, Toshiki Yagi, and Ritsu Kamiya. Direct measurement of inter-doublet elasticity in flagellar axonemes. Cell structure and function, 24(1):27–33, 1999. 24. B. Alberts. Molecular Biology of the Cell. Garland Science, New York, 2002. 25. J. Howard. Mechanics of motor proteins and the cytoskeleton. Sinauer Associates Sunderland, MA, 2001. . . . . . . . . . . . . . . . . . . . . . . . 12 r s o s . r o y a s o c e i l t y p u b l i i s h n g . o r g . . . . . . . . . . . . . R . . . . . S o c . . . . . . . o p e n . . s c i . . . . . . . . . . . . . 0 0 0 0 0 0 0 . . 26. Charles B. Lindemann, Walter G. Rudd, and Robert Rikmenspoel. The stiffness of the flagella of impaled bull sperm. Biophysical Journal, 13(5):437 – 448, 1973. 27. M. Okuno. Inhibition and relaxation of sea urchin sperm flagella by vanadate. The Journal of Cell Biology, 85(3):712, 1980. 28. M. Okuno and Y. Hiramoto. Direct measurements of the stiffness of echinoderm sperm flagella. Journal of Experimental Biology, 79(1):235, 1979. 29. SS Antman. Nonlinear Problems of Elasticity, volume 107 of Applied Mathematical Sciences. Springer, 2005. 30. Charles B. Lindemann, Lisa J. Macauley, and Kathleen A. Lesich. The counterbend phenomenon in dynein-disabled rat sperm flagella and what it reveals about the interdoublet elasticity. Biophysical Journal, 89(2):1165 –1174, 2005. 31. Dominic W Pelle, Charles J Brokaw, Kathleen A Lesich, and Charles B Lindemann. Mechanical properties of the passive sea urchin sperm flagellum. Cell Motil Cytoskeleton, 66(9):721–735, Sep 2009. 32. H. Gadêlha, E. A. Gaffney, and A. Goriely. The counterbend phenomenon in flagellar axonemes and cross-linked filament bundles. Proceedings of the National Academy of Sciences, July 2013. 33. Gang Xu, Kate S Wilson, Ruth J Okamoto, Jin-Yu Shao, Susan K Dutcher, and Philip V Bayly. Flexural rigidity and shear stiffness of flagella estimated from induced bends and counterbends. Biophysical Journal, 110(12):2759–2768, 2016. 34. R. Everaers, R. Bundschuh, and K. Kremer. Fluctuations and stiffness of double-stranded polymers: railway-track model. EPL (Europhysics Letters), 29:263, 1995. 35. M. Hines and JJ. Blum. Bend propagation in flagella. i. derivation of equations of motion and their simulation. J. Biophys., 23:41, 1978. 36. Charles J. Brokaw. 37. Charles J. Brokaw. Molecular mechanism for oscillation in flagella and muscle. Proc. Natl Acad. Sci., 72:3102, 1975. Computer simulation of flagellar movement. vi. simple curvature-controlled models are incompletely specified. J. Biophys., 48:633, 1985. 38. C.J. Brokaw and D.R. Rintala. Computer simulation of flagellar movement. iii. models incorporating cross-bridge kinetics. J. Mechanochem. Cell. Motil., 3(2):77, 1975. 39. K. E. Machin. The control and synchronization of flagellar movement. Proc. R. Soc. London, Ser. B, 158:88, 1963. 40. Charles B Lindemann. A" geometric clutch" hypothesis to explain oscillations of the axoneme of cilia and flagella. Journal of theoretical biology, 168(2):175–189, 1994. 41. M Hines and JJ Blum. Bend propagation in flagella. ii. incorporation of dynein cross-bridge kinetics into the equations of motion. Biophysical journal, 25(3):421, 1979. 42. Robert Rikmenspoel. Contractile mechanisms in flagella. Biophysical journal, 11(5):446, 1971. 43. Ingmar H. Riedel-Kruse, Andreas Hilfinger, Jonathon Howard, and Frank Jülicher. How molecular motors shape the flagellar beat. HFSP Journal, 1(3):192–208, 2007. 44. Philip V. Bayly and Kate S. Wilson. Equations of Interdoublet Separation during Flagella Motion Reveal Mechanisms of Wave Propagation and Instability. Biophysical Journal, 107(7):1756–1772, October 2014. 45. P. V. Bayly and K. S. Wilson. Analysis of unstable modes distinguishes mathematical models of flagellar motion. Journal of The Royal Society Interface, 12(106):20150124, May 2015. 46. Pablo Sartori, Veikko F. Geyer, Andre Scholich, Frank Jülicher, and Jonathon Howard. Dynamic curvature regulation accounts for the symmetric and asymmetric beats of Chlamydomonas flagella. eLife, 5:e13258, May 2016. 47. Charles J Brokaw. Computer simulation of flagellar movement x: doublet pair splitting and bend propagation modeled using stochastic dynein kinetics. Cytoskeleton, 71(4):273–284, 2014. 48. Charles B Lindemann and Kathleen A Lesich. Flagellar and ciliary beating: the proven and the possible. Journal of Cell Science, 123(4):519–528, 2010. 49. Charles J Brokaw. Computer simulation of flagellar movement ix. oscillation and symmetry breaking in a model for short flagella and nodal cilia. Cell motility and the cytoskeleton, 60(1):35–47, 2005. 50. C.J. Brokaw and D.R. Rintala. Computer simulation of flagellar movement. iii. models incorporating cross-bridge kinetics. J. Mechanochem. Cell. Motil., 3(2):77, 1975. 51. David Oriola, Hermes Gadêlha, and Jaume Casademunt. Nonlinear amplitude dynamics in flagellar beating. Royal Society Open Science. 52. Franck Plouraboue, Ibrahima Thiam, Blaise Delmotte, Eric Climent, PSC Collaboration, et al. Identification of internal properties of fibers and micro-swimmers. In APS Meeting Abstracts, 2016. 53. A. Hilfinger, A. K. Chattopadhyay, and F. Jülicher. 13 . . . . . . . . . . . . . . . . . . . . . . . . . . r s o s . r o y a s o c e i l t y p u b l i i s h n g . o r g . . . . . . . . . . R . S o c . . . . . . . . . . . o p e n . . . . . . s c i . . . . . . . 0 0 0 0 0 0 0 . . . . Nonlinear dynamics of cilia and flagella. Phys. Rev. E, 79:051918–051925, 2009. 54. E.M. Purcell. Life at low reynolds number. Am. J. Phys., 45:3, 1977. 55. Francesco Pampaloni, Gianluca Lattanzi, Alexandr Jonáš, Thomas Surrey, Erwin Frey, and Ernst-Ludwig Florin. Thermal fluctuations of grafted microtubules provide evidence of a length-dependent persistence length. Proceedings of the National Academy of Sciences, 103(27):10248–10253, 2006. 56. Hermes Gadêlha. On the optimal shape of magnetic swimmers. Regular and Chaotic Dynamics, 18(1-2):75–84, 2013. 57. Kirsty Y Wan and Raymond E Goldstein. Coordinated beating of algal flagella is mediated by basal coupling. Proceedings of the National Academy of Sciences, 113(20):E2784–E2793, 2016. 58. E.A. Gaffney, H. Gadêlha, D.J. Smith, J.R. Blake, and J.C. Kirkman-Brown. Mammalian sperm motility: Observation and theory. Annual Review of Fluid Mechanics, 43(1):501–528, 2011. 59. Raymond E. Goldstein, Marco Polin, and Idan Tuval. Emergence of synchronized beating during the regrowth of eukaryotic flagella. Phys. Rev. Lett., 107:148103, Sep 2011. 60. Greta Quaranta, Marie-Eve Aubin-Tam, and Daniel Tam. Hydrodynamics versus intracellular coupling in the synchronization of eukaryotic flagella. Phys. Rev. Lett., 115:238101, Nov 2015. 61. David Oriola, Hermes Gadêlha, C Blanch-Mercader, and Jaume Casademunt. 62. James D Murray. Subharmonic oscillations of collective molecular motors. EPL (Europhysics Letters), 107(1):18002, 2014. Mathematical Biology. II Spatial Models and Biomedical Applications {Interdisciplinary Applied Mathematics V. 18}. Springer-Verlag New York Incorporated, 2001. Spatial patterns: higher order models in physics and mechanics, volume 45. Springer Science & Business Media, 2012. 65. Arkady Pikovsky, Michael Rosenblum, Jürgen Kurths, and Robert C Hilborn. Synchronization: a universal concept in nonlinear science. American Journal of Physics, 70(6):655–655, 2002. 63. Donald S Cohen and James D Murray. A generalized diffusion model for growth and dispersal in a population. Journal of Mathematical Biology, 12(2):237–249, 1981. 64. Lambertus A Peletier and William C Troy. 14 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . r s o s . r o y a s o c e l i t y p u b l i i s h n g . o r g . . . . . . . . . . . . . R . S o c . . . . . o p e n s c i . . . . . . . . . 0 0 0 0 0 0 0 . . . . . . . .
1206.5946
1
1206
2012-06-26T10:44:38
Entanglement and Sources of Magnetic Anisotropy in Radical Pair-Based Avian Magnetoreceptors
[ "physics.bio-ph", "physics.chem-ph" ]
One of the principal models of magnetic sensing in migratory birds rests on the quantum spin-dynamics of transient radical pairs created photochemically in ocular cryptochrome proteins. We consider here the role of electron spin entanglement and coherence in determining the sensitivity of a radical pair-based geomagnetic compass and the origins of the directional response. It emerges that the anisotropy of radical pairs formed from spin-polarized molecular triplets could form the basis of a more sensitive compass sensor than one founded on the conventional hyperfine-anisotropy model. This property offers new and more flexible opportunities for the design of biologically inspired magnetic compass sensors.
physics.bio-ph
physics
Entanglement and Sources of Magnetic Anisotropy in Radical Pair-Based Avian Magnetoreceptors Hannah J. Hogben, Till Biskup, and P. J. Hore∗ Department of Chemistry, University of Oxford, Physical & Theoretical Chemistry Laboratory, Oxford, OX1 3QZ, UK. (Dated: May 14, 2018) One of the principal models of magnetic sensing in migratory birds rests on the quantum spin- dynamics of transient radical pairs created photochemically in ocular cryptochrome proteins. We consider here the role of electron spin entanglement and coherence in determining the sensitivity of a radical pair-based geomagnetic compass and the origins of the directional response. It emerges that the anisotropy of radical pairs formed from spin-polarized molecular triplets could form the basis of a more sensitive compass sensor than one founded on the conventional hyperfine-anisotropy model. This property offers new and more flexible opportunities for the design of biologically inspired magnetic compass sensors. PACS numbers: 03.67.Bg, 82.30.Cf, 87.50.C- The biophysics and biochemistry that allow birds to sense the direction of the geomagnetic field (25-65 µT) are for the most part obscure. One of the two cur- rently popular hypotheses (the other involves biogenic iron-oxide nanostructures [1]) is founded on magneti- cally sensitive photochemical reactions in the retina [2]. It is thought that photo-induced radical pairs in cryp- tochrome, a blue-light photoreceptor protein, may con- stitute the primary magnetic sensor [3, 4] and a variety of supporting evidence has accumulated over the last few years (reviewed in [5–8]). If this mechanism proves to be correct, it will incontrovertibly come under the umbrella of 'quantum biology' [9], as an instance of Nature us- ing fundamentally quantum behaviour – in this case the coherent spin dynamics of radical pairs – to achieve some- thing that would be essentially impossible by means of more conventional chemistry. For this reason, the avian magnetic compass has attracted the attention of quan- tum information theorists and others wishing to under- stand the role played by spin-entanglement and to deter- mine whether the techniques of quantum control could shed light on this intriguing sensory mechanism [10–13]. A fundamental property of radical pairs that allows sensitivity to magnetic interactions orders of magnitude smaller than kBT is that their chemical transformations conserve electron spin. Radical pairs are therefore cre- ated with the same spin-multiplicity (singlet or triplet) as their precursors. Owing to electron-nuclear hyper- fine (HF) interactions, neither singlets nor triplets are, in general, eigenstates of the spin Hamiltonian. Con- sequently, the radical pair starts out in a non-stationary superposition which evolves coherently at frequencies de- termined by the HF interactions and also, crucially for a magnetic sensor, by the electronic Zeeman interactions with an external magnetic field [5]. Spin decoherence and spin relaxation can be slow enough to allow even an Earth-strength magnetic field to modulate the spin dy- namics and hence alter the yields of the products formed by spin-selective reactions. The anisotropy of the HF in- teractions leads to anisotropic reaction yields and hence, in principle, a magnetic direction sensor [14, 15]. The singlet state – the initial state of the radical pairs formed photochemically in cryptochromes [4, 16] – is en- tangled: 2α1β2(cid:105)(cid:104)α1β2 + 1 S(cid:105)(cid:104)S = 1 − 1 2α1β2(cid:105)(cid:104)β1α2 − 1 2β1α2(cid:105)(cid:104)β1α2 2β1α2(cid:105)(cid:104)α1β2 (1) (α and β are the mS = ± 1 2 spin states of the two unpaired electrons). But other initial states are also known to re- sult in magnetically sensitive chemistry [17]: do they too need to be entangled or is it sufficient if they are 'merely' coherent? Or is neither entanglement nor coherence nec- essary for a magnetic compass? Questions such as these have been addressed in two re- cent papers. Briegel and his group noted that randomly generated separable (i.e. not entangled) initial states could result in reaction product yields more anisotropic than those produced from an initial singlet state under the same conditions [10]. The other study, by Benjamin and colleagues, reached similar conclusions by analysing model radical pair systems, finding significant product yield anisotropies for the separable initial state [11] 1 2S(cid:105)(cid:104)S + 1 2α1β2(cid:105)(cid:104)α1β2 2T0(cid:105)(cid:104)T0 = 1 2β1α2(cid:105)(cid:104)β1α2 + 1 (2) in which T0 is the mS = 0 triplet spin state. Here we examine the role of initial entanglement and attempt to clarify the various sources of magnetic anisotropy that might form the basis of a radical pair compass sensor in birds. Initial radical pair states. We start by identifying chemically feasible initial electron spin states. Gemi- nate radical pairs are normally formed by spin-conserving chemical reactions so that at the moment of their creation they are either pure singlet, described by the initial elec- tron spin density matrix ρ0 = ρ0(S) = S(cid:105)(cid:104)S, or pure triplet ρ0 = ρ0(T) = 1 3 (cid:0)11 − S(cid:105)(cid:104)S(cid:1). Occasionally, singlet 2 1 0 2 n u J 6 2 ] h p - o i b . s c i s y h p [ 1 v 6 4 9 5 . 6 0 2 1 : v i X r a and triplet formation channels operate in parallel [18], in which case ρ0 is a weighted sum of ρ0(S) and ρ0(T), i.e. of S(cid:105)(cid:104)S and 11: ρ0 = µρ0(S) + (1 − µ)ρ0(T) = 1 3 (4µ − 1)S(cid:105)(cid:104)S + 1 3 (1 − µ)11 (3) Eq. (3) is also appropriate for 'F-pairs' [17] formed from radicals with uncorrelated spins (i.e. µ = 1 4 ). The op- erators S(cid:105)(cid:104)S and 11 and their linear combinations are invariant to rotations in the electron spin-space, mean- ing that all states that can be written in the form of Eq. (3) are isotropic. Any ρ0 that cannot be so written is necessarily anisotropic. Significantly different initial states can occur when the radical pair comes from a molecular triplet precursor formed by intersystem crossing (ISC). This route is com- mon in photochemical reactions of the general type: AB hν−−→ S[AB]∗ ISC−−−→ T[AB]∗ reaction −−−−−−→ T[A• B•] (4) in which the final step that creates the triplet radical pair could be homolysis (as shown) or inter- or intramolecu- lar electron transfer, hydrogen atom transfer, etc. The formation of T[AB]∗ from S[AB]∗ requires the creation of spin angular momentum at the expense of orbital angular momentum. This process is mediated by spin-orbit cou- pling and is anisotropic in the molecular frame [19]. That is, the three triplet sub-levels of T[AB]∗ are differentially populated leading to a spin polarization in the molecular frame that is passed to the radical pair on its formation. In an appropriately chosen molecular axis system, the initial state of the radical pair may be written: (cid:88) ρ0 = pqTq(cid:105)(cid:104)Tq (5) q=x,y,z Anisotropic ISC is known to be responsible for a variety of spin-chemical and spin-polarization phenomena [17, 20–22]. Aside from linear combinations of Eqs (3) and (5), there are no other commonly occurring initial conditions for radical pairs subject to weak magnetic fields. Minimal radical pair model. Insights into the spin dy- namics of the various initial states just identified can be obtained from a minimal model [23] comprising two elec- tron spins one of which is coupled to a spin-1⁄2 nucleus (e.g. 1H). The HF interaction is either isotropic or axi- ally anisotropic according to the value of a dimensionless parameter, α [14]. Two cases are considered specifically: α = 0 (isotropic) and α = −1 (the anisotropic interac- tion that results in the largest reaction yield anisotropy for this 3-spin system [13]). To account for the chemi- cal reactivity of the radical pair, we adopt the 'exponen- tial model' [23] in which singlet and triplet states react spin-selectively with the same first-order rate constant, k, to form distinct products. The quantum yields of these competing reactions are calculated using standard 2 methods [14, 23] (outlined in the Appendix). The two quantities of interest are ΦS, the fractional yield of the product formed via the singlet pathway, referred to here as the 'reaction yield', and ∆ΦS, the magnitude of its anisotropy: ∆ΦS = max{ΦS}− min{ΦS}. The variation of ΦS with the orientation of the radical pair in a 50 µT magnetic field is the basis of the compass sensor. To begin, we choose the isotropic initial condition in Eq. (3) together with an anisotropic HF interaction (α = −1). In the not unrealistic limit, a (cid:29) ω (cid:29) k [14, 24]: 124µ − 1 (6) 12 (4µ − 1) cos2 θ; ∆ΦS = 1 ΦS = 1 4 + 1 where a is the isotropic HF coupling constant, ω is the strength of the magnetic field, and θ is the angle between the symmetry axis of the HF tensor and the magnetic field vector. ΦS is anisotropic, and therefore potentially suitable as a magnetic compass, except when the initial state is a statistical (1⁄4 : 3⁄4) mixture of singlet and triplet (µ = 1 4 ) occurs when the initial state is pure singlet (µ = 1); for a pure triplet initial state (µ = 0), ∆ΦS is smaller by a factor of three. These results were verified by exact numerical simulations (see Appendix). 4 ). The maximum anisotropy (∆ΦS = 1 To quantify the entanglement of the various initial elec- tron spin states considered here, we use the 'concurrence' C(ρ0) proposed by Wootters [25] for a two-qubit density operator. For the initial condition in Eq. (3), C(ρ0) is 2µ− 1 when µ > 1 2 (see Appendix). Thus, a singlet–triplet mixture must contain more than 50% singlet for the initial state to be entangled. The pure triplet state (µ = 0) is not entangled, but as we have just seen it gives rise to a significantly anisotropic reaction yield. 2 and zero when µ ≤ 1 We now turn to a different initial condition, a linear combination of Eq. (3) (with µ = 0) and Eq. (5) (with px = py = 0; pz = 1): ρ0 = ηS(cid:105)(cid:104)S + (1 − η)Tz(cid:105)(cid:104)Tz (7) i.e. an anisotropic mixed singlet-triplet initial state in which the triplet component is 100% polarized along the molecular z-axis. In the same limit as before (a (cid:29) ω (cid:29) k), but now for an isotropic HF interaction: ΦS = 3 8 − 1 4 (1 − η) sin2 θ; ∆ΦS = 1 4 (1 − η) (8) where θ is now the angle between the triplet polarization axis (z) and the magnetic field vector. The anisotropy is maximised when η = 0 (pure Tz(cid:105) triplet, ∆ΦS = 1 4 ) and is at a minimum when η = 1 (pure singlet, ∆ΦS = 0). Once again, these expressions were confirmed by numer- ical simulations (see Appendix). We note that Eqs (6) and (8) predict identical maximum directional responses. The reaction yield is isotropic when η = 1 because then both the initial state S(cid:105)(cid:104)S and the spin-Hamiltonian are isotropic. The angle-dependence in Eq. (8) clearly arises because the spin dynamics depend on the direction of the magnetic field with respect to the quantization (z) axis of the initial Tz(cid:105) state [26]. The concurrence of the density operator in Eq. (7) is 2η− 1 when η ≥ 1 2 and 1− 2η when 2 . Pure singlet and pure Tz(cid:105) triplet thus have the η ≤ 1 same degree of entanglement but lead to very different ∆ΦS. 3 4 ρ0(S) + 3 Hitherto we have taken the reaction rates of the singlet and triplet states (kS and kT) to be identical. Once this restriction is lifted, it is even possible to have magnetic field effects when the initial state is a statistical mixture 11. To of singlet and triplet: ρ0 = 1 illustrate this point, simulations for the minimal radical pair with an anisotropic HF coupling are included in the Appendix. ∆ΦS is non-zero except when kS = kT. That is, a radical pair can exhibit magnetic compass proper- ties even when its initial electron spin state is neither entangled nor coherent. In this case the coherence arises during the spin evolution as a result of the differential reactivity of the singlet and triplet states. 4 ρ0(T) = 1 4 Relation between compass properties and entangle- ment. A complex picture emerges from these simple considerations. Entangled initial states can give small or zero reaction yield anisotropy. Non-entangled ini- tial states can lead to appreciable anisotropy. With two sources of anisotropic reaction yields – the initial state and the HF interactions – it is tricky to assess whether entanglement, or coherence in a given basis, is essen- tial for magnetic compass action. For example, replac- 2T0(cid:105)(cid:104)T0 ing ρ0 = S(cid:105)(cid:104)S (Eq. (1)) by ρ0 = 1 (Eq. (2)) not only removes the initial entanglement, and the coherence in the {α1β2(cid:105),β1α2(cid:105)} basis, it also intro- duces anisotropy that was not present in S(cid:105)(cid:104)S. Simi- larly, most randomly chosen initial states are anisotropic and some will give a larger ∆ΦS than does S(cid:105)(cid:104)S under identical conditions. In short, it appears that initial en- tanglement is not a particularly helpful concept when as- sessing the sensitivity of a radical pair compass; nor is it straightforwardly illuminating to consider the behaviour of artificial initial states. 2S(cid:105)(cid:104)S + 1 pair compass based above suggest A radical the largest considerations on initial-state anisotropy. The an alternative compass design in which the directionality comes from the initial condition rather than the HF interactions. In the minimal model, the initial state that gives reaction yield anisotropy is ρ0 = Tq(cid:105)(cid:104)Tq where q = x, y, z (see Appendix). We therefore compare Tq(cid:105)(cid:104)Tq with S(cid:105)(cid:104)S using exact numerical simulations (see Appendix). The possibility that spin-polarized triplet radical pairs might offer some advantage over singlets has been noted before but without realistic suggestions for the chemical origin of such initial states [26]. Figure 1 shows the reaction yield anisotropy of a radi- cal pair inspired by the flavin adenine dinucleotide radi- cal, FADH•, formed photochemically in cryptochromes [27]. One radical contains 1H and 14N nuclei with Reaction yield anisotropy, ∆ΦS, calculated (see FIG. 1. Appendix) for a radical pair in which one radical contains a 1H nucleus (spin-1⁄2) and a 14N nucleus (spin-1). k = 106 s−1 and ω = 50 µT. The HF coupling parameters (in mT) are: aH = −0.8; TH,xx = 0.8 δ; TH,yy = −0.6 δ; TH,zz = −0.2 δ; aN = 0.4; TN,xx = −0.5 δ; TN,yy = −0.5 δ; TN,zz = 1.0 δ. ρ0 = S(cid:105)(cid:104)S (black) and ρ0 = Ty(cid:105)(cid:104)Ty (green). Also shown are representations of the hyperfine tensors for δ = 0 (left) and δ = 1 (right). isotropic HF couplings approximately equal to those of the proton and nitrogen (H5 and N5, see appendix) in the central ring of the tricyclic isoalloxazine ring system of FADH• (these being the two largest HF interactions in FADH• [28]). The anisotropic components of the two interactions were also modelled on FADH•, but with a uniform scaling by a factor of δ, in the range 0.001− 1.0. For the smaller values of δ, the spin-Hamiltonian is es- sentially isotropic. When the initial state ρ0 is a 100% spin-polarized triplet, ∆ΦS has significant magnitude for In contrast, when ρ0 = S(cid:105)(cid:104)S, ∆ΦS is all values of δ. essentially zero until the HF tensors become significantly anisotropic (δ ≈ 0.1). By the time the HF anisotropy is comparable to that in FADH• (i.e. δ ≈ 1.0), both ini- tial states give very similar directional responses to the 50 µT applied magnetic field. This suggests that a spin- polarized triplet geminate radical pair with isotropic HF interactions could operate as a compass sensor just as well as an initial singlet state with anisotropic HF inter- actions. Indeed, there are circumstances in which, other things being equal, the anisotropy of the initial state might offer a more sensitive compass than one based on HF anisotropy. Biologically plausible radical pairs are likely to have many magnetic nuclei (mostly 1H and 14N) with differently aligned HF tensors. Simulations suggest that the directional information potentially available from in- dividual HF tensors tends to be scrambled in a multi- nuclear radical pair, resulting in a greatly reduced ∆ΦS (see Appendix). A simple illustration of this effect is given in Fig. 2 which shows simulations of the reaction yield anisotropy for a spin system in which one of the rad- -3-2-100.000.050.100.15NHNH 4 plant [4] and frog [16]), flavin-tryptophan radical pairs are formed as singlets. However, avian cryptochromes may behave differently, and there are precedents for triplet radical pairs in other flavoproteins [30, 31]. Su- perficially, it appears that flavins may be suitable for an initial triplet-state compass: intersystem crossing in both flavin mononucleotide and riboflavin at near-neutral pH results in fractional populations of the zero-field triplet sub-levels of px = 1 3 , pz = 0 [32]. Within the min- imal model discussed above, this would lead to a high re- action yield anisotropy, two-thirds that of the maximum possible (see Appendix). 3 , py = 2 The use of spin-polarized triplets should open new channels for the design of bio-inspired molecular devices for sensing the direction of weak magnetic fields. We thank DARPA (QuBE: N66001-10-1-4061) and the EPSRC for financial support. Appendix Basis states. The spin dynamics of radical pairs may usefully be described in terms of two distinct sets of basis states. In both, singlet and triplet states are eigenstates of the total electron spin operator, S: (cid:104)S SS(cid:105) = 0 √ (cid:104)Ti STi(cid:105) = 2 (i = 0,±1 or x, y, z) (9) The triplet basis states are either the eigenstates of Sz, the component of S along the z-axis: (cid:104)Tm SzTm(cid:105) = m (m = 0,±1) (10) or are defined in terms of the three cartesian components of S: (cid:104)Tq SqTq(cid:105) = 0 (cid:104)Tx SyTz(cid:105) = i (q = x, y, z) (11) (and cyclic permutations of x, y, z) The relations between the two are: T−1(cid:105) − 1√ T−1(cid:105) + i√ Tx(cid:105) = 1√ Ty(cid:105) = i√ Tz(cid:105) = T0(cid:105) 2 2 2 2 S(cid:105) and Tm(cid:105) can also be written: T+1(cid:105) T+1(cid:105) S(cid:105) = 1√ [α1β2(cid:105) − β1α2(cid:105)] 2 T+1(cid:105) = α1α2(cid:105) T0(cid:105) = 1√ T−1(cid:105) = β1β2(cid:105) 2 [α1β2(cid:105) + β1α2(cid:105)] where αj(cid:105) and βj(cid:105) are defined by: (cid:104)αj Sj,zαj(cid:105) = + 1 (cid:104)βj Sj,zβj(cid:105) = − 1 2 2 (j = 1, 2) (12) (13) (14) FIG. 2. Reaction yield anisotropy, ∆ΦS, calculated (see Appendix) for a radical pair in which one radical contains four 1H nuclei, all of which have axially anisotropic HF in- teractions with a = 0. The symmetry axes of the four HF tensors are directed towards the vertices of a tetrahedron. Three of the tensors have principal values: T11 = T22 = −1.0, T33 = 2.0 mT. The fourth is identical apart from a uniform scaling of the principal values by a factor δ. k = 106 s−1 and ω = 50 µT. ρ0 = S(cid:105)(cid:104)S (black) and ρ0 = Tx(cid:105)(cid:104)Tx (green). Also shown are representations of the hyperfine tensors for δ = 0.5, δ = 1.0, and δ = 1.5. icals contains four spin-1⁄2 nuclei with tetrahedrally dis- posed axial HF tensors. When all four tensors are identi- cal (δ = 1), the reaction yield anisotropy for ρ0 = S(cid:105)(cid:104)S vanishes, by symmetry. However, when the symmetry is reduced to C3v, by scaling the principal components of one of the HF tensors by a factor δ, the value of ∆ΦS increases but does not approach that afforded by ρ0 = Tx(cid:105)(cid:104)Tx until log10 δ reaches ca. 1.0. Thus it appears that the compass properties of a radical pair with many mutually cancelling HF interactions could be 'rescued' by having a triplet, rather than a singlet, ini- tial condition, provided the triplet is spin-polarized by anisotropic intersystem crossing. Discussion. Having identified the initial spin-states in which radical pairs may be formed by chemical reac- tion, we revisited earlier attempts to determine the im- portance of entanglement and coherence as determinants of the anisotropic responses of radical pair magnetorecep- tors. It appears that the use of artificial initial spin-states for this purpose is somewhat confounded by their intrin- sic anisotropy, the effects of which may dominate the anisotropy conferred by the HF interactions. From these considerations it emerges that the anisotropy of radical pairs formed from spin-polarized molecular triplets could form the basis for a magnetic compass that is more sen- sitive than one based on the conventional HF-anisotropy model [2] in particular when the HF couplings are not strongly anisotropic or when the individual effects of mul- tiple HF anisotropies tend to counteract one another. Would a triplet radical pair compass be compatible with cryptochrome as the primary magnetoreceptor? In the cryptochromes investigated hitherto (bacterial [29], -1.00.01.0-0.50.50.000.020.040.060.08 The axis system in which Sx, Sy and Sz and Sj,z are defined may be chosen to be the 'laboratory frame', in which the z-axis is commonly the direction of the applied magnetic field, or a 'molecular frame', which could, for example, be the principal axis system of one of the hy- perfine (HF) tensors. The triplet state Tq(cid:105) (q = x, y, z), Eq. (11), is spin-polarized in the q = 0 principal plane within the molecule [19]. 5 Initial state. The initial state of the radical pair spin system, ρ(0), is written as the direct product of the initial density operator for the two electron spins, ρ0, and iden- tity operators for each of the nuclear spins (i = 1, 2,··· ) to which the electrons are coupled: (cid:41) (cid:40)(cid:79) i ρ(0) = ρ0 ⊗ 1 M 11i (15) (M is the total dimension of the nuclear spin-space). It is assumed that the formation of the radical pair is not nuclear spin-dependent. In the absence of chemical re- activity and spin-decoherence, the probability that the radical pair is in a singlet state at time t is given by the expectation value of the singlet projection operator, P S: (cid:68) P S(cid:69) (t) = Tr = Tr (cid:104) (cid:104) ρ(t) P S(cid:105) e−i Ht ρ(0)e+i Ht P S(cid:105) (cid:41) (cid:40)(cid:79) 11i i P S = {S(cid:105)(cid:104)S} ⊗ (16) (17) where H is the time-independent spin Hamiltonian and Hhfi = (cid:88) (cid:88) q=x,y,z = a Minimal model. The principles of radical pair magne- toreception can be discussed using a simple model com- prising two electron spins and one spin-1⁄2 nucleus (e.g. 1H) with an axially anisotropic HF coupling to one of the electron spins: Aqq S1,q Iq S1,q Iq + aα (cid:16) S1,x Ix + S1,y Iy − 2 S1,z Iz (cid:17) (18) q=x,y,z a is the isotropic part of the HF interaction (expressed as an angular frequency) and α is a dimensionless axiality parameter [14]. Defined in this way, the HF interaction has cylindrical symmetry around the molecular z-axis. Two cases are considered specifically: α = 0 (isotropic HF interaction) and α = −1 (the HF interaction that results in the largest anisotropy in the reaction yield of this 3-spin system [13]). The electron Zeeman interaction is included by means of the spin Hamiltonian: (cid:88) (cid:104) Sj,z cos θ + Sj,x sin θ (cid:105) (19) HZeeman = ω j=1,2 Reaction yield anisotropy, ∆ΦS, of the minimal FIG. 3. 11. radical pair model with a non-coherent initial state, ρ0 = 1 4 ∆ΦS is shown as a function of the rate constants kS and kT. a = 1.0 mT, α = −1, ω = 50 µT. in which ω is the strength of the applied magnetic field (expressed as an angular frequency) and θ specifies its direction with respect to the symmetry axis (z) of the HF tensor. It is assumed that the g-tensors of the two radicals are identical and isotropic, and that the nuclear Zeeman interactions are negligible. Both are excellent approximations for organic radicals subject to the weak magnetic fields of interest here. To account for the chemical reactivity of the radical pair within the minimal model, we use the 'exponential model' [23] in which singlet and triplet states react spin- selectively with the same first-order rate constant, kS = kT = k, to form distinct products. Although unlikely to be strictly valid for any real magnetoreceptor, this approximation simplifies the algebra without distorting the underlying physics. The yield of the chemical product formed via the singlet pathway, and its anisotropy, are calculated as [23]: (cid:90) ∞ (cid:68) P S(cid:69) (t)e−ktdt ΦS = k ∆ΦS = max (ΦS) − min (ΦS) 0 (21) so that 0 ≤ ΦS ≤ 1. The corresponding yield for the triplet reaction channel is simply 1 − ΦS. ΦS is referred to as the reaction yield and ∆ΦS as the reaction yield anisotropy. The variation of ΦS with the orientation of the radical pair with respect to an external magnetic field forms the basis of the compass mechanism [2]. When kS (cid:54)= kT, the calculation of ΦS is performed in Liouville space: ΦS = kS L = i (cid:68) P S L−1ρ(0) (cid:69) (cid:16) H ⊗ 118 − 118 ⊗ H T(cid:17) (cid:16) P S ⊗ 118 + 118 ⊗ P S(cid:17) (cid:16) P T ⊗ 118 + 118 ⊗ P T(cid:17) 2 kS + 1 + 1 2 kT (20) (22) (23) 4567890.00.10.20.30.40.5 6 initial electron spin state is neither entangled nor coher- ent. The coherence arises during the spin evolution as a result of the differential reactivity of the singlet and triplet states. Perturbation theory. To obtain estimates of the max- imum possible magnetic responses within the minimal model, we use a perturbative approach [14, 24], appropri- ate for weak applied magnetic fields and long-lived radical pairs. This approximation is valid when a (cid:29) ω (cid:29) k. These conditions are not unrealistic: HF interactions are of the order of 108 rad s−1 (≈ 500 µT), the geomagnetic field is roughly 107 rad s−1 (≈ 50 µT), and plausible values of the rate constant k are 105 − 106 s−1 [5]. Eqs (6) and (8) were verified by the exact numerical simulations, the results of which are shown in Fig. 4. Multinuclear radical pairs. In the general case, the HF component of the spin Hamiltonian has the form: (cid:88) (cid:88) j=1,2 (cid:88) (cid:88) k (cid:105) (cid:104)Sj · Ajk · Ik (cid:104) j=1,2 k Hhfi = = ajk Sj · Ik + Sj · Tjk · Ik (25) (cid:105) where ajk, Tjk and Ajk are, respectively, the isotropic HF coupling constant, the anisotropic HF tensor and the total HF tensor for nucleus k coupled to the electron in radical j. The Zeeman term is: HZeeman = ω (cid:88) (cid:104) Sj,x sin θ cos φ j=1,2 + Sj,y sin θ sin φ + Sj,z cos θ (26) (cid:105) where θ and φ specify the direction of the field in the molecular axis system. Fig. 5 shows, for completeness, versions of Fig. 1 in which the three initial triplet states are compared with S(cid:105)(cid:104)S. Concurrence. To quantify the entanglement of the various initial electron spin states ρ0, we use the 'con- currence' proposed by Wootters for a two-qubit density operator [25]: C(ρ0) = max{0, λ1 − λ2 − λ3 − λ4} (27) where the λi are the non-negative real square roots of the eigenvalues, in decreasing order, of: ρ0 (σy ⊗ σy) ρ∗ 0 (σy ⊗ σy) (28) in which σy is twice the Sy operator for a single electron spin, and ρ∗ 0 is the complex conjugate of ρ0. initial General conditions. Within the minimal model, the most general initial state consistent with both Eqs (3) and (5) is: (cid:88) q=x,y,z FIG. 4. Magnetic field effect on the reaction yield of the minimal radical pair model. k = 10−3a. (a) Anisotropic HF interaction (α = −1). S (solid lines) and T (dashed lines) denote the initial radical pair states ρ0 = ρ0(S) (µ = 1) and ρ0 = ρ0(T) (µ = 0), respectively, in Eq. (3). The pertur- bation theory result in Eq. (6) is valid in the shaded region where a (cid:29) ω (cid:29) k. The dependence of ΦS on the strength of the applied magnetic field (ω/a) is shown for various angles between the symmetry axis of the HF tensor and the mag- netic field vector. The sharp features near log10(ω/a) = 0.3 arise from level anti-crossings [24]. (b) Isotropic HF inter- action (α = 0). S (solid line) and T (dashed lines) denote ρ0 = S(cid:105)(cid:104)S (η = 1) and ρ0 = Tz(cid:105)(cid:104)Tz (η = 0), respectively, in Eq. (7). The perturbation theory result in Eq. (8) is valid in the shaded region where a (cid:29) ω (cid:29) k. The dependence of ΦS on the strength of the applied magnetic field (ω/a) is shown for various angles between the triplet alignment axis and the magnetic field vector. where 118 is the identity operator in the 8-dimensional spin-space and P T = 118 − P S. Fig. 3 shows simulations for the minimal radical pair with an anisotropic HF coupling and an initial state: ρ0 = 1 4 ρ0(S) + 3 4 ρ0(T) = 1 4 11 (24) ∆ΦS is non-zero except when kS = kT. The radical pair can exhibit magnetic compass properties even when its ρ0 = εS(cid:105)(cid:104)S + (1 − ε) pqTq(cid:105)(cid:104)Tq (29) (a)S, 0°S, 30°S, 45°S, 60°S, 90°T, 0°T, 30°T, 45°T, 60°T, 90°0.30.20.1-4-202(b)-4-2020.70.60.50.40.30.20.10.0ST, 0°T, 30°T, 45°T, 60°T, 90° TABLE I. HF data for N5 and H5 in FADH•. 7 N5 0.393 Principal axes Nucleus a / mT Tjj / mT −0.498 −0.492 0.990 −0.0384 −0.616 0.9819 −0.168 −0.0348 0.784 −0.1861 0.8655 −0.2432 0.4380 0.8981 −0.4097 0.1595 0.2883 0.9568 0.1883 −0.0203 0.2850 0.9579 0.9398 −0.2864 −0.769 H5 = max{px, py, pz} pmax = where min{px, py, pz}. The maximum anisotropy is ob- tained when ε = 0, pmax = 1 and pmin = 0 (giving ∆ΦS = 1 pmin and 4 ). The concurrence C(ρ0) of the state in Eq. (29) is: 2ε − 1 when ε > 1 2pmax(1 − ε) − 1 when ε ≤ 1 2 2 and pmax ≥ 1 2(1−ε) 0 otherwise (32) (33) (34) Laboratory-frame polarization Although not relevant for a geomagnetic compass sensor, radical pairs can be created with large laboratory-frame polarizations, i.e. with unequal populations of the triplet eigenstates in a strong magnetic field (Tm(cid:105) (m = 0,±1), as defined by Eq. (10), with the z-axis being the direction of a strong applied magnetic field). For example, the Triplet Mecha- nism of Chemically Induced Dynamic Electron Polariza- tion can result in large polarizations for radical pairs pro- duced by triplet states formed by anisotropic intersystem crossing [20, 33]. Another example, which also requires the electron spins to be quantized by strong electron Zee- man interactions, is seen in the EPR spectra of spin- correlated radical pairs with non-zero electron-electron exchange and/or dipolar interactions [34, 35]. Intersystem crossing. The implications of anisotropic intersystem crossing for the behaviour and properties of radical pairs is discussed in detail by Steiner and Ulrich [17] (pp. 109–112). In most cases, as here, intersystem crossing is assumed to be independent of HF-coupled nuclear spins. However, Kothe et al. [22], in an ele- gant study of quantum oscillations in an organic triplet state, have shown that the nuclei are in fact involved and that the appropriate molecular-frame triplet basis states are eigenstates of the combined zero-field and hyperfine Hamiltonians. We do not consider this possibility here. FIG. 5. Reaction yield anisotropy, ∆ΦS, calculated for a radical pair in which one radical contains a 1H nucleus (spin- 1⁄2) and a 14N nucleus (spin-1). k = 106 s−1 and ω = 50 µT. The HF coupling parameters (in mT) are as in the caption for Fig. 1. ρ0 = S(cid:105)(cid:104)S (black). ρ0 = Tx(cid:105)(cid:104)Tx (a, green), ρ0 = Ty(cid:105)(cid:104)Ty (b, green), ρ0 = Tz(cid:105)(cid:104)Tz (c, green). with 0 ≤ ε ≤ 1, px, py, pz ≥ 0 and px + py + pz = 1. Some such states are entangled and some are not. Almost all are anisotropic and may lead to compass behaviour even when the spin-Hamiltonian is isotropic. For the minimal model, with an isotropic HF coupling and a (cid:29) ω (cid:29) k, the initial state in Eq. (29) leads to: ΦS = 1 8 (2ε + 1) + 1 (cid:0)px sin2 θ cos2 φ + py sin2 θ sin2 φ + pz cos2 θ(cid:1) 4 (1 − ε) (30) ∆ΦS = 1 4 (1 − ε)(pmax − pmin) (31) -3-2-100.000.050.100.15-3-2-100.000.050.100.15(b)(c)-3-2-100.000.050.100.15(a)NRNNHNOOH5 FADH• radical. The HF interactions for FADH• used to calculate the reaction yield anisotropies shown in Fig. 2 are based on the data in Table I [28]. ∗ [email protected] [1] M. Winklhofer and J. L. Kirschvink, J. R. Soc. Interface 7, S273 (2010). [2] K. Schulten, C. E. Swenberg, and A. Weller, Z. Phys. Chemie 111, 1 (1978). [3] T. Ritz, S. Adem, and K. Schulten, Biophys. J. 78, 707 (2000). [4] K. Maeda, A. J. Robinson, K. B. Henbest, H. J. Hogben, T. Biskup, M. Ahmad, E. Schleicher, S. Weber, C. R. Timmel, and P. J. Hore, Proc. Natl. Acad. Sci. USA 109, 4774 (2012). [5] C. T. Rodgers and P. J. Hore, Proc. Natl. Acad. Sci. USA 106, 353 (2009). [6] M. Liedvogel and H. Mouritsen, J. R. Soc. Interface 7, S147 (2010). [7] T. Ritz, Procedia Chem. 3, 262 (2011). [8] H. Mouritsen and P. J. Hore, Curr. Opin. Neurobiol. 22, 343 (2012). [9] P. Ball, Nature 474, 272 (2011). [10] J. Cai, G. G. Guerreschi, and H. J. Briegel, Phys. Rev. Lett. 104, 220502 (2010). [11] E. M. Gauger, E. Rieper, J. J. L. Morton, S. C. Benjamin, and V. Vedral, Phys. Rev. Lett. 106, 040503 (2011). [12] C. Y. Cai, Q. Ai, H. T. Quan, and C. P. Sun, Phys. Rev. A 85, 022315 (2012). [13] J. Cai, F. Caruso, and M. B. Plenio, Phys. Rev. A 85, 040304 (2012). [14] F. Cintolesi, T. Ritz, C. W. M. Kay, C. R. Timmel, and P. J. Hore, Chem. Phys. 294, 385 (2003). [15] K. Maeda, K. B. Henbest, F. Cintolesi, I. Kuprov, C. T. Rodgers, P. A. Liddell, D. Gust, C. R. Timmel, and P. J. Hore, Nature 453, 387 (2008). 8 [16] S. Weber, T. Biskup, A. Okafuji, A. R. Marino, T. Berthold, G. Link, K. Hitomi, E. D. Getzoff, E. Schle- icher, and J. R. Norris, Jr., J. Phys. Chem. B 114, 14745 (2010). [17] U. E. Steiner and T. Ulrich, Chem. Rev. 89, 51 (1989). [18] K. Maeda, C. J. Wedge, J. G. Storey, K. B. Henbest, P. A. Liddell, G. Kodis, D. Gust, P. J. Hore, and C. R. Timmel, Chem. Commun. 47, 6563 (2011). [19] M. S. de Groot, I. A. M. Hesselmann, and J. H. van der Waals, Mol. Phys. 12, 259 (1967). [20] P. W. Atkins and G. T. Evans, Mol. Phys. 27, 1633 (1974). [22] G. Kothe, T. Yago, [21] A. Katsuki, Y. Kobori, S. Tero-Kubota, S. Milikisyants, H. Paul, and U. E. Steiner, Mol. Phys. 100, 1245 (2002). J.-U. Weidner, G. Link, M. Lukaschek, and T.-S. Lin, J. Phys. Chem. B 114, 14755 (2010). [23] C. R. Timmel, U. Till, B. Brocklehurst, K. A. McLauch- lan, and P. J. Hore, Mol. Phys. 95, 71 (1998). [24] C. R. Timmel, F. Cintolesi, B. Brocklehurst, and P. J. Hore, Chem. Phys. Lett. 334, 387 (2001). [25] W. K. Wootters, Phys. Rev. Lett. 80, 2245 (1998). [26] G. E. Katsoprinakis, A. T. Dellis, and I. K. Kominis, New J. Phys. 12, 085016 (2010). [27] T. Langenbacher, D. Immeln, B. Dick, and T. Kottke, J. Am. Chem. Soc. 131, 14274 (2009). [28] S. Weber, K. Mobius, G. Richter, and C. W. M. Kay, J. Am. Chem. Soc. 123, 3790 (2001). [29] T. Biskup, K. Hitomi, E. D. Getzoff, S. Krapf, and S. Weber, Angew. T. Koslowski, E. Schleicher, Chem. Int. Ed. 50, 12647 (2011). [30] W. Eisenreich, M. Joshi, S. Weber, A. Bacher, M. Fischer, J. Am. Chem. Soc. 130, 13544 (2008). and [31] S. S. Thamarath, J. Heberle, P. J. Hore, T. Kottke, and J. Matysik, J. Am. Chem. Soc. 132, 15542 (2010). [32] R. M. Kowalczyk, E. Schleicher, R. Bittl, and S. Weber, J. Am. Chem. Soc. 126, 11393 (2004). [33] P. J. Hore, Chem. Phys. Lett. 69, 563 (1980). [34] C. D. Buckley, D. A. Hunter, P. J. Hore, and K. A. McLauchlan, Chem. Phys. Lett. 135, 307 (1987). [35] P. J. Hore, D. A. Hunter, C. D. McKie, and A. J. Hoff, Chem. Phys. Lett. 137, 495 (1987).
1009.5101
2
1009
2010-11-23T15:51:23
Nematic order by elastic interactions and cellular rigidity sensing
[ "physics.bio-ph" ]
We predict spontaneous nematic order in an ensemble of active force generators with elastic interactions as a minimal model for early nematic alignment of short stress fibers in non-motile, adhered cells. Mean-field theory is formally equivalent to Maier-Saupe theory for a nematic liquid. However, the elastic interactions are long-ranged (and thus depend on cell shape and matrix elasticity) and originate in cell activity. Depending on the density of force generators, we find two regimes of cellular rigidity sensing for which orientational, nematic order of stress fibers depends on matrix rigidity either in a step-like manner or with a maximum at an optimal rigidity.
physics.bio-ph
physics
Nematic order by elastic interactions and cellular rigidity sensing B. M. Friedrich, S. A. Safran Department of Materials and Interfaces, Weizmann Institute of Science, 76100 Rehovot, Israel (Dated: November 23, 2010) Abstract We predict spontaneous nematic order in an ensemble of active force generators with elastic inter- actions as a minimal model for early nematic alignment of short stress fibers in non-motile, adhered cells. Mean-field theory is formally equivalent to Maier-Saupe theory for a nematic liquid. However, the elastic interactions are long-ranged (and thus depend on cell shape and matrix elasticity) and originate in cell activity. Depending on the density of force generators, we find two regimes of cellular rigidity sensing for which orientational, nematic order of stress fibers depends on matrix rigidity either in a step-like manner or with a maximum at an optimal rigidity. PACS numbers: 61.30.Gd, 87.10.Pq, 87.16.Ln Keywords: active force dipole, acto-myosin cytoskeleton, stress fiber, Eshelby theory of elastic inclusions, ferro- elasticity 0 1 0 2 v o N 3 2 ] h p - o i b . s c i s y h p [ 2 v 1 0 1 5 . 9 0 0 1 : v i X r a 1 I. INTRODUCTION The actin cytoskeleton of living cells comprises semiflexible actin filaments and force- generating myosin molecular motors (as well as many regulatory proteins) [1]. In different cell types, the actin cytoskeleton can display quite distinct degrees of order: In many motile cell types, the bulk cytoskeleton is simply a cross-linked filament meshwork without any higher degree of ordering, while in striated muscle cells, actin and myosin assemble into crystal-like myofibril aggregates [2]. We are interested in an intermediate state of cytoskeletal order found in non-motile, adhered cells, where crosslinked bundles of parallelly aligned actin filaments form, so called stress fibers [3]. These stress fibers often align with the long axis of the cell, which establishes nematic order within the ensemble of stress fibers and results in polarized cell forces. Interestingly, experiments demonstrate a strong influence of substrate rigidity on the nematic ordering of stress fibers. In recent experiments with stem cells, the degree of nematic order of nascent stress fibers showed a non-monotonic dependence on substrate rigidity with a maximum at an optimal rigidity [4]. More generally, substrate rigidity was shown to be a determining factor for morphology and could even trigger cell fate decisions [5]. In this paper, we address the question of why nascent stress fibers align preferentially with the long axis of the cell in a substrate dependent manner. Our work provides a unified theoretical foundation for a previous study, which addressed this question in the framework of a linear response theory that treated the entire cell as a single elastic inclusion [4]. We predict more general phase behavior for nematic ordering as a function of substrate rigidity, cell shape anisotropy and contractile force strength. We focus on the early stages of symmetry breaking by considering a minimal model of force generators that is motivated by short, nascent stress fibers, which are coupled to the remaining, disordered cytoskeleton and which exert contractile forces. The resulting elastic deformations of the cytoskeleton provide a mean of long-range communication between distant parts of the cell [6]. We propose that such elastic interactions between nascent stress fibers drive their nematic alignment with respect to the long axis of the cell in a way that depends on cell shape and substrate rigidity. More precisely, we study the emergence of nematic order within an ensemble of active force generators. We employ a generic description of the force generators in terms of extended force dipoles, which are subject to mutual elastic interactions [7, 8]. The force generators are ho- mogenously distributed within the cellular domain for which we assume suitable boundary conditions. For simplicity, we model the disordered cytoskeleton as an isotropic, linear elas- tic material. While the assumption of isotropy might be well justified for the early stages of cytoskeletal symmetry breaking, the assumption of a linear elastic material is a strong sim- plification: The cytoskeleton is known to be viscoelastic on long time-scales and may exhibit non-linear elastic behavior for large stresses. Despite these limitations, our minimal model pro- vides qualitative insight into cytoskeletal polarization and reveals different regimes of rigidity sensing. We find that the cellular shape determines a macroscopic strain field due to the long-ranged nature of elastic interactions. It thus provides a cue for nematic ordering similar to the shape- dependent depolarization factors in electrostatics [9]. However, in contrast to the electrostatic case where the shape modifies the response to an external electric field, the elastic case of ac- 2 tive force dipoles shows shape-dependent nematic order even in the absence of external stresses [10]. Additionally, generic, isotropic hard-core repulsion between force generators can favor cooperative nematic alignment within the ensemble. Using a mean field approach inspired by the classical theory of orientational polarization of interacting dipoles [9], we derive the de- pendence of nematic ordering on macroscopic boundary conditions (such as domain shape and rigidity of a surrounding matrix) from pairwise elastic interactions. This dependence provides our prototypical cell with a mechanism to sense and respond to substrate stiffness [11]. We use a mean field approach to predict nematic order of nascent stress fibers as a function of physical parameters of the surrounding matrix. In addition to a regime of non-monotonic dependence of nematic order on matrix stiffness with a maximum at some optimal rigidity, which had been already found in [4], we find a second regime, which is characterized by a monotonic, step-like dependence. Experiments on cells with different aspect ratios observed qualitatively different polarization responses that match the two regimes of our theoretical model. Our approach thus generalizes and unifies previous work that either described active cellular responses by a phe- nomenological linear response theory [4], or considered elastic interactions between individual point force dipoles, but without reference to cell shape or matrix rigidity [12]. II. EXTENDED FORCE DIPOLES We consider a simple model of a mesoscopic force generator embedded in an elastic material, which we propose as a generic description of the tension forces exerted by a short stress fiber on the surrounding, disordered cytoskeleton. A general force generator located at x = 0 exerts a field fj(x) of active forces on the cytoskeleton and induces a strain field uij(x). The force field fj(x) is characterized by its multipole moments Pi1,...,in,j =R d3x xi1 · · · xinfj(x) [8]. Note that the monopole moment vanishes, Pi = 0, since the force generator is not acted upon by external forces. Thus the dipole moment Pij dominates the far field strain. The far-field strain at distances x much larger than the size a of the force generator equals to leading order the strain field ufar induced by a point force dipole with localized force dipole density ij pij(x) = Pij δ(x) [8]. If the cell were infinite, uij ≈ ufar ij = Gik,jl(x/x, νc) Pkl/(Ecx3) (1) where Ec is the Young's modulus of the cell and G is the angular part of the force dipole Green's function that depends on its Poisson ratio νc [8, 13]. Note that the strain field is long-ranged and decays as x−3 just like the electric field of an electric dipole, but the symmetry of this field expressed by G is akin to that of an electric quadrupole. The field ufar ij does not accurately reflect the strain field at small distances comparable to the size a of the force generator. In particular, ufar ij diverges at x = 0. Here, we abstract from fine details of the strain field uij at small distances and propose to employ a as a cut-off distance by introducing a force dipole density pij(x) that takes the constant value Pij/B inside a spherical region B = {x < a} of volume B = 4πa3/3, but is zero outside B. Note that the force dipole density pij(x) is equivalent to a set of surface forces efi = pijnj acting on the boundary of B, where nj is the surface normal of the region B. This is analogous to the situation in electrostatics, where a uniform electric dipole density induces the same electrical field as a set of surface charges [9]. 3 Using Eshelby's classical theory of elastic inclusions, we find that the strain field uin spherical region B is uniform [14, 15]. Its purely dilational strain uin,0 isotropic part p0 (which does not involve any volume change) is linear in the anisotropic part ps ij = pkk δij/3 of the dipole density, while the pure shear strain uin,s ij = uin ij inside the kkδij/3 is linear in the ij − uin,0 ij = uin ij = pij − p0 ij ij uin,0 ij = α0p0 ij/Ec and uin,s ij = αsps ij/Ec. (2) The coefficients α0 = (1/3)(1 + νc)/(1 − νc)(1 − 2νc) and αs = (2/15)(4 − 5νc)/(1 − νc)(1 + νc) are positive for all physical values of the Poisson ratio −1 < νc < 1/2 [14]. For incompressible matrices characterized by νc = 1/2, the coefficient α0 vanishes. The strain field uout ij outside the region B is equivalent to a strain field induced by the fictitious point force dipole Pijδ(x) not require the explicit form of the outer strain field; for later calculations, it is sufficient to over a shell concentric with B vanishes. The inner and outer strain field are related by a jump condition that involves the fictitious surfaces force plus some fictitious point force octupole ePijklδ(x) located at the center of the ball [15]. We will note that its spherical mean Rx=c d2x uout,k ef mentioned above, uin ij (an) = 3α0 efinj/Ec for n = 1 [15]. Previous work considered point force dipoles, which correspond to the limit a → 0. Our use of extended force dipoles avoids singularities, in particular the elastic deformation energy associated with a force dipole is finite. Additionally, hard-core repulsion between different force dipoles is introduced in a natural way. The extended force dipole provides a prototypical example of a force generator and will be used throughout this paper. ij − uout ij j i n(k) ij = P0 n(k) . The individual dipole moments P (k) We now consider an ensemble of homogenously distributed, extended force dipoles of density ρ that reside inside an elastic domain Ω. The dipoles have random positions xk and dipole moments P (k) are bipolar with director nk. ij Such a dipole moment equals the dipole moment of a pair of opposing point forces ±P0/a nk separated by a vector 2ank. The elastic domain Ω and the force dipoles mimic the contractile activity of short, nascent stress fibers embedded into disordered cytoskeleton within an non- motile, adhered cell. We assume that the domain Ω is surrounded by an infinitely extended, elastic matrix that represents the physical substrate to which the cell adheres, see fig. 1. The Young's modulus Em and the Poisson ratio νm of this matrix may differ from the respective values Ec and νc of the cellular domain Ω. To capture the effect of an anisotropic domain geometry, yet keep the calculations simple, we consider spheroidal domains with semi-axis ax = ay < az and aspect ratio r = az/ax. Thus the z-axis denotes the axis of revolution of the cellular domain. We treat the dipole strength P0 as a control parameter; experiments suggest that P0 may increase with matrix stiffness [16]. We assume that the directors nk of the individual force dipoles are free to rotate (while their positions xk are fixed for simplicity). In our particular case of cellular domain that has the shape of a prolate spheroid, we can show that polarization only occurs along the axis of revolution of the cellular domain. We can therefore characterize nematic order by a single, scalar order parameter S = (3/2)hcos2 θkik − (1/2), cos θk = nk,z. (3) This order parameter S takes the maximal value 1 for perfect alignment with the z-axis and 4 E , ν m m x y E , ν c c Ω z FIG. 1: We present a minimal model for the elastic interactions between short, nascent stress fibers within a non-motile, adhered cell. The short stress fibers are modeled as an ensemble of force generators that reside inside an elastic cellular domain Ω with Young's modulus Ec and Poisson ratio νc. The cellular domain in turn is embedded in an elastic matrix (with Young's modulus Em and Poisson ratio νm) that mimics the substrate to which the cell adheres. vanishes for an isotropic director distribution. For later use, we also introduce the alignment parameters Sk = (3/2) cos2 θk − (1/2) of all the individual dipoles. III. ELASTIC INTERACTIONS R d3x (Σlu(l) The elastic deformation energy due to the ensemble of extended force dipoles is H = is the elastic stress field induced by the k-th dipole [13]. Using the fact that the force due to an ensemble of dipoles is proportional to the gradient of the dipole density (∇iσ(k) ij )/2 where σ(k) ij )(Σkσ(k) ij ij = ∇ip(k) ij ), we rewrite H as [8] H =Xk,l Ukl, Ukl = 1 2Zx−xk<a d3s u(l) ij p(k) ij . (4) Here, Ukk represents a "self-energy" of the k-th dipole[24], whereas Ukl represents a pairwise elastic interaction energy which can be written in the form Ukl ∼ P (k) It has been proposed that active force generators that are fueled by an external energy reservoir minimize the work needed to deform the matrix while maintaining a constant force magnitude (here P0/a) [17][25]. This minimization principle successfully describes the behavior of whole cells adhered to an elastic substrate in a phenomenological way. Assuming that cytoskeletal reorganization is governed by local mechanosensitive feedback mechanisms, we argue that this minimization principle also applies to sub-cellular cytoskeletal structures such as nascent stress fibers. This is a crucial assumption of our work and can be tested by comparing our theory to future experiments. ij Gim,jnP (l) mn [7, 8]. 5 IV. MACROSCOPIC STRAIN FIELDS DEPEND ON BOUNDARY CONDITIONS We introduce the macroscopic strain field uij that averages the microscopic strain field Σku(k) ij on a length-scale that is much larger than both the dipole ball size a and the typical dipole-dipole distance ρ−1/3. Due to the linear relationship between the strain field and the stress sources, this macroscopic strain field is equivalent to the strain field induced by the macroscopic force dipole density pij = ρP ij where P ij = hP (k) ij i is the ensemble average of the force dipole moments [26]. The force dipole ensemble generates an isotropic macroscopic dipole density p0 ij = pkkδij/3 = ρP0δij/3 irrespective of nematic ordering. This "hydrostatic pressure" has no analogue in the electrostatic case of electric dipoles, which have polar as opposed to tensor, nematic symmetry. A net polarization of the dipole ensemble (here assumed in the z-direction) gives an additional anisotropic part ps ij = ρP0S(δizδjz − δij/3) which is proportional to the nematic order parameter S. For domains of spheroid shape, the corresponding macroscopic strain is homogenous throughout the domain Ω [14], similar to the situation in electrostatics [9]. This homogenous strain uij can be split into a dilation part u0 ij = ukkδij and a pure shear part us ij. Only the shear exerts a torque on a force dipole and is therefore the quantity of interest to us. From symmetry considerations, we infer that the shear strain can be written in the form ij = uij − u0 ij = pij − p0 us ij = (−h + gS) ρP0 (δizδjz − δij/3)/Ec. (5) The coefficients h and g can again be computed using Eshelby's theory of elastic inclusions [14]; they are depicted in fig. 1(a) as a function of matrix rigidity Em for different values of the domain aspect ratio r. The coefficient g/Ec couples the z-anisotropic part of the macroscopic force dipole density to the macroscopic shear strain us ij and thus provides a macroscopic analogue of the shear compliance 1/(2µc) = (1 + νc)/Ec. The coefficient h is novel in the elastic case and does not exist in the simple electrostatic case; it couples a "hydrostatic pressure" p0 ij to a shear strain; this shape field factor must vanish by symmetry in the case of a spherical domain, but may be non-zero for non-spherical domains due an anisotropic distribution of elastic restoring forces from the matrix. Below, we infer the qualitative behavior of h and g by discussing four important limit cases: (i) The case of a matrix that is much stiffer than the cellular domain with Em ≫ Ec corresponds to clamped boundary conditions. Hence, the macroscopic elastic strain uij vanishes within the cellular domain and g = h = 0. In this case, restoring forces from the matrix completely counterbalance the active cell forces. (ii) If the matrix is much softer than the cellular domain, Em ≪ Ec, the elastic domain Ω has essentially stress-free boundary conditions, and there are no restoring forces from the matrix. Hence, the macroscopic elastic stress σij within Ω equals pij. From ps ij, we conclude g = 2µcEc = 1 + νc and h = 0, i.e. there is no shape-dependence of the macroscopic strain field. (iii) If Ω is a sphere s and has the same elastic properties as the matrix, then us ij/Ec is computed analogous to the inner strain field of an extended force dipole. Hence, g = αs, h = 0. (iv) Finally, in the limiting case of a prolate spheroid with r → ∞, i.e. an infinite rod, we have uzz = 0, and hence h = g > 0. For a proof, consider the special case of a dipole density that is fully polarized in z-direction with S = 1. The strain uij can be equivalently induced by a set of surface forces ij = 2µcus ij = αs ρP ij = σs efi = pijnj. For a cylinder, efi = 0 as nz = 0. Thus, uij = 0 for S = 1, which implies h − g = 0. 6 h (shape field factor) g (domain shear compliance) 1+ν r=1 1 r=4 r=∞ 0 0 0.02 0.2 α s E /E m c 1 2 10 20 ∞ 0.1 0 r=∞ r=4 r=1 r=∞ r=4 r=1 E /E m c 0 0.02 0.2 1 2 10 20 ∞ FIG. 2: The macroscopic force dipole density inside the cellular domain (pij) induces shear strain within this domain (us ij) in two different ways, which are characterized by coefficients g and h, see eqn. (5). Both coefficients depend on the normalized stiffness Em/Ec of the matrix surrounding the domain. This is shown for the case of a perfectly spherical cellular domain with r = 1, as well as for the case of prolate spheroid with aspect ratio r = 4 and for the case of an infinite cylinder (r = ∞). The "domain shear compliance" g characterizes the amount of shear strain, which is due to an anisotropy of the force dipole density itself. The "shape field factor" h characterizes the shear strain that is induced as a consequence of an asymmetric domain shape and anisotropic restoring forces from the surrounding matrix by the isotropic "hydrostatic pressure" part of the force dipole density. For the figure, the Poisson ratios are νc = νm = 0.3; the limiting cases discussed in the text are marked in red. It can be shown that h > 0 holds true for any prolate spheroid with 0 < r ≤ ∞. Consistent with the limiting cases discussed above, h is a non-monotonic function of Em/Ec that vanishes if the matrix is either very soft or very stiff. The domain shear compliance g is a monotonically decreasing function of matrix rigidity Em since the restoring forces, which oppose shear strain us ij, increase with Em. ij = Pl6=k u(l) The local strain field in the vicinity of a typical dipole at position xk induced by all the other dipoles, uloc,k ij , is the superposition of the macroscopic strain field uij and a local correction term, which is similar to a Lorentz cavity field [9] and insensitive to boundary conditions [27]. In a continuum approximation that averages over the random positions xl of the other dipoles [28], the local strain is induced by a homogeneous dipole density pij that fills the entire cellular domain, except for a spherical void around xk because of hard-core repulsion. Thus, the local correction approximately amounts to subtracting the strain contribution of a dipole density pij filling the void and we obtain for the shear strain hhuloc,k,s ij (x)iik = us ij − αsps ij/Ec for x − xk < a, (6) By construction, the averaged local strain field looks the same for all force dipoles, except possible those which are very close to the boundary of the domain Ω. 7 (a) Monotonic order-response for strong interaction, high density, low temperature 1 S 0.5 r=4 r=1 1 S 0 1 10 J/T 10 2 E /E m c 0 0 0.02 0.1 0.2 1 2 10 20 ∞ (b) Non-monotonic order-response for weak interaction, low density, high temperature 0.2 S 0.1 0 r=4 ν =.3c ν =.45 c r=1 E /E m c 0 0.02 0.1 0.2 1 2 10 20 ∞ FIG. 3: Two regimes of rigidity sensing. The nematic order parameter S of the force dipole ensemble depends strongly on the normalized stiffness Em/Ec of the matrix. Dependent on the effective strength J/T ∗ of their elastic interaction, we distinguish two regimes for this dependence: (a) We observe a monotonic, step-like dependence on Em/Ec for strong interactions with J ≫ T ∗, i.e. for strong dipole strength P0, high dipole density ρ and low effective temperature T ∗. A similar dependence is found for S as a function of J/T ∗ if Em is fixed (here Em = 2Ec), see inset. (b) We observe a non-monotonic dependence with a maximum of nematic order at an optimal matrix stiffness for weak interactions with J ≪ T ∗, i.e. for weak dipole strength P0, low dipole density ρ and high effective temperature T ∗. The Poisson ratios are always νc = νm = 0.3, except for panel (b), where we show in gray also the case νc = 0.45. V. MEAN FIELD THEORY FOR NEMATIC ORDER The Hamiltonian H = Pk,l Ukl of pairwise elastic interactions of the force dipoles can be rewritten as a coupling between their dipole moments and the respective local strain fields (plus a self-energy term that is independent of the dipole orientations) H = 1 2Xk Zx−xk<a d3x uloc,k ij P (k) ij /B + Ukk. (7) We derive a mean-field Hamiltonian H0 of Lorentz-Weiss type from this Hamiltonian by replac- ing the true local strain fields uloc,k iik. Using eq. (5) (and by their mean field averages hhuloc,k ij ij 8 a standard variational technique, which results in a prefactor 2 [18]), we find Zx−xk<a H0 =Xk = const. −Xk d3x hhuloc,k ij iikP (k) ij /B J ( h −egS ) Sk (8) with eg = g − αs. Here J = 2ρP 2 0 /(3Ec) sets the energy scale of elastic interactions. This energy scale should be compared to an effective temperature T ∗ that quantifies the noise in the system [19]. Assuming for simplicity Boltzmann statistics for the dipole orientations, we self-consistently solve for the nematic order parameter S [20]. In the limit of low interaction or negative as the domain shear compliance g > 0 and the cavity effect of isotropic hard-core strength egJ ≪ T ∗, we find S ≈ hJ/(5T ∗ +egJ). For prolate spheroids, the shape field with h is always positive and favors nematic order. The coefficient eg = g − αs can be either positive repulsion (αs) give competing contributions to eg. From eqn. (6), this can be understood in terms of increased or decreased structural interference of the strain induced by the "central" dipole at xk with the macroscopic shear strain us ij/Ec, respectively. ij and the local cavity correction −αsps The elastic interaction energy of eq. (8) is formally equivalent to the generic theory of nematic liquid crystals of Maier and Saupe [21] with an additional, external alignment field [20]. In our case, this field arises from the contractility of the system (which has no analogue in the electrostatic case) and does not require an external stress. The field strength hJ is proportional to the shape field factor h and thus depends on the shape of the cellular domain as well as the rigidity of the matrix. Recall that for h = 0 (i.e. for spherical domains or very stiff or soft matrices), Maier-Saupe theory predicts a first order phase transition from an isotropic to a nematic phase when egJ/T ∗ ≈ −4.542 [21]. A necessary condition for this transition is eg = g − αs < 0, i.e. the cavity effect of isotropic hard-core repulsion has to outweigh the domain shear compliance g. In the special case of a spherical domain with r = 1, the shape field vanishes, h = 0, and we have g < αs for stiff matrices with Em > Ec, see fig. 2. Accordingly, we find strong nematic order for Em ∼> Ec, see fig. 3(a). The full phase diagram exhibits a line of first order phase transition that ends in a critical point, see [20] and fig. 4. Generally, positive values of h increase nematic order and smooth out the sharp phase transition that exists when h = 0, see fig. 3(a) for the case r = 4. This monotonic, step-like dependence of the nematic order parameter S on matrix rigidity Em is found whenever the elastic interactions are sufficiently strong with αsJ/T ∗ > 4.542. For weak elastic interactions (or large noise strength), the nematic order parameter S emulates the behavior of the shape field factor h, and hence depends non-monotonically on matrix rigidity Em for prolate spheroids, see figs. 2 and 3(b). If the cellular domain Ω is a perfect sphere with r = 1 and elastic interactions are weak, there is no nematic order. The results shown are robust with respect to changes of the Poisson ratio νm of the the matrix, and, in the regime of a monotonic nematic response, also to changes in the Poisson ratio νc of the cellular domain (not shown). Only the non-monotonic nematic response is attenuated in nearly incompressible cellular domains and vanishes for νc = 1/2, see fig. 3(b). In conclusion, the nematic order of force generators provides a read-out of matrix rigidity and exhibits qualitatively different regimes of dependency for strong and weak elastic interactions. 9 FIG. 4: Phase diagram for the nematic order parameter S of an ensemble of force dipoles as a function of the coefficients of the Hamiltonian H0, see eqn. (8). In our cell model, the coefficientseg and h depend in turn on domain shape and matrix stiffness in a non-trivial way, see fig. 2. As illustration of this dependence, we traced out these coefficients for an example parameter set for a prolate domain (r = 4, νc = νm = 0.3), which correspond to the black Em-S curve in fig. 3(b). It should be noted that our mean field theory is based on a Weiss approximation and does temperature. In a real cell, anisotropic hard-core repulsion of actin bundles can favor nematic not correct for a reaction field [22]; we thus underestimate eg and overestimate the transition order [23] and thus effectively decrease eg. Both corrections stem from local effects and are insensitive to boundary conditions. A non-monotonic dependence of cell force polarization on matrix rigidity was already pre- dicted in [4] in the framework of a linear response theory, which assumed ps ij. This linear response theory corresponds to the regime of weak elastic interaction J ≪ T ∗ in our theory. Our theory allows a self-consistent derivation of the phenomenological parameter χs, which represents an orientational polarization susceptibility: In the limit J ≪ T ∗, we find χs = (Ec/5)J/T ∗. Other work studied nematic order by elastic interactions within a two- dimensional ensemble of force dipoles using Monte-Carlo simulations and found nematic order for high dipole densities [12]. The use of periodic boundary conditions in that study is equivalent to a strain-free boundary and should be compared to our limit Em ≫ Ec. ij = −χsus VI. CONCLUSION We showed that elastic interactions can drive nematic ordering of cytoskeletal force gener- ators such as short, nascent stress fibers as a function of matrix rigidity. A non-monotonic dependence of actin cytoskeleton polarity on matrix rigidity has indeed been found in experi- ments with embryonic stem cells [4], which corresponds to our limit of a low density of force generators and large noise strength. Interestingly, in high aspect ratio cells, nematic order sat- urates at high matrix stiffness [4]; this may correspond to our prediction of saturation at high 10 values of J/T ∗ since these cells display more developed stress fibers and hence higher contractile forces. We speculate that different cells might employ different mechanisms of rigidity sensing by controlling e.g. the strength of active cell forces. VII. ACKNOWLEDGEMENT We thank D. Discher, N. Gov, F. Rehfeldt, A. Zemel for stimulating discussions; this work was supported by the Israel Science Foundation, the Schmidt Minerva Center (SAS), the Ger- man Academic Exchange Service (BMF) and the historic generosity of the Perlman Family Foundation. [1] J. Howard. Mechanics of Motor Proteins and the Cytoskeleton. Sinauer, 2001. [2] B. Alberts, D. Bray, J. Lewis, M. Raff, K. Roberts, and J. D. Watson. Molecular Biology of the Cell. Garland Publ. Inc., third edition, 1994. [3] S. Deguchi and M. Sato. Biomechanical properties of actin stress fibers of non-motile cells. Biorheology, 46:93 -- 105, 2009. [4] A. Zemel, D. Discher, and S. Safran. Optimal matrix rigidity for stress-fibre polarization in stem cells. Nat. Physics, 6:468 -- 473, 2010. [5] A. J. Engler, S. Sen, H. L. Sweeney, and D. E. Discher. Matrix elasticity directs stem cell lineage specification. Cell, 126:677, 2006. [6] J. Akst. Full speed ahead: Physical forces acting in and around cells are fast -- and making waves in the world of molecular biology. The Scientist, 23:26, 2009. [7] H. Wagner and H. Horner. Elastic interactions and the phase transition in coherent metal- hydrogen systems. Adv. Phys., 23:587 -- 637, 1974. [8] U. S. Schwarz and S. A. Safran. Elastic interactions of cells. Phys. Rev. Lett., 88(4):048102, 2002. [9] C. Kittel. Introduction to Solid State Physics. John Wiley and Sons, New York, 1986. [10] A. A. Zemel and S. A. Safran. Active self-polarization of contractile cells in asymmetrically shaped domains. Phys. Rev. E, 76(2):021905, 2007. [11] D. E. Discher, P. Janmey, and Y.-L. Wang. Tissue cells feel and respond to the stiffness of their substrate. Science, 18:1139 -- 1143, 2005. [12] I. B. Bischofs and U. S. Schwarz. Effect of Poisson ratio on cellular structure formation. Phys. Rev. Lett., 95:068102, 2005. [13] L. D. Landau, L. P. Pitaevskii, and E. M. Lifshitz. Theory of elasticity. Butterworth-Heinemann, New York, 1986. [14] J. D. Eshelby. The determination of the elastic field of an ellipsoidal inclusion, and related problems. Proc. R. Soc. London, Ser. A, 241:376, 1957. [15] T. Mura. Micromechanics of Defects in Solids. Kluwer Academic, Dordrecht, 1991. [16] A. Zemel, F. Rehfeldt, A. E. X. Brown, D. E. Discher, and S. A. Safran. Cell shape, spreading symmetry and the polarization of stress-fibers in cells. J. Phys.: Condens. Matter, 22:194110, 2010. 11 [17] I. B. Bischofs and U. S. Schwarz. Cell organization in soft media due to active mechanosensing. Proc. Natl. Acad. Sci., 100:9274 -- 9279, 2003. [18] S. A. Safran. Statistical Thermodynamics of Surfaces, Interfaces and Membranes. Westview, 2003. [19] A. Zemel, I. B. Bischofs, and S. A. Safran. Active elasticity of gels with contractile cells. Phys. Rev. Lett., 97(12):128103, 2006. [20] P. J. Wojtowicz and P. Sheng. Critical point in the magnetic field-temperature phase diagram of nematic liquid crystals. Phys. Rev. Lett., 48:235 -- 236, 1974. [21] W. Maier and A. Saupe. Eine einfache molekular-statistische Theorie der nematischen kristallinflussigen Phase. Teil I. Z. Naturforschg., 14:882 -- 889, 1959. [22] H. Thomas and R. Brout. Molecular field theory, the onsager reaction field, and the spherical model. J. appl. Phys., 39:624, 1965. [23] L. Onsager. Ann. N. Y. Acad. Sci., 51:627, 1949. [24] The self-energy Ukk diverges, if the spatial cut-off a → 0; a UV-cut-off in Fourier space yields a regularized Ukk similar to ours, T. Lubensky (private communication). [25] This is different from passive dipoles, for which their own energetics has to be included [7]. [26] Unlike [4], we choose the reference state uij = 0 for zero activity of the force generators with pij = 0. [27] Boundary conditions affect the local correction as terms of order R−2, R is the distance between xk and domain boundary; L. Walpole, Int. J. Eng. Sci. 34, 629 (1996). [28] For regular dipole arrangements [12], crystal fields can favor nematic order (i.e. for a sc lattice), but they can also favor anti-ferroelastic order (i.e. for a fcc lattice). 12
1308.1128
1
1308
2013-08-05T21:43:10
Embedding a carbon nanotube across the diameter of a solid state nanopore
[ "physics.bio-ph", "cond-mat.mes-hall", "cond-mat.mtrl-sci", "cond-mat.soft", "physics.ins-det" ]
A fabrication method for positioning and embedding a single-walled carbon nanotube (SWNT) across the diameter of a solid state nanopore is presented. Chemical vapor deposition (CVD) is used to grow SWNTs over arrays of focused ion beam (FIB) milled pores in a thin silicon nitride membrane. This typically yields at least one pore whose diameter is centrally crossed by a SWNT. The final diameter of the FIB pore is adjusted to create a nanopore of any desired diameter by atomic layer deposition (ALD), simultaneously embedding and insulating the SWNT everywhere but in the region that crosses the diameter of the final nanopore, where it remains pristine and bare. This nanotube-articulated nanopore is an important step towards the realization of a new type of detector for biomolecule sensing and electronic characterization, including DNA sequencing.
physics.bio-ph
physics
Embedding a carbon nanotube across the diameter of a solid state nanopore Running title: Embedding a carbon nanotube across a nanopore Running Authors: Sadki et al. E.S. Sadkia), S. Garaj, D. Vlassarev, J.A. Golovchenko, and D. Branton b) Harvard University, Department of Physics, 17 Oxford Street, Cambridge, MA 02138 USA a)Present address: Physics Department, Faculty of Science, United Arab Emirates University, Al Ain, P.O. Box 17551, Abu Dhabi, UAE b)Harvard University, Department of Molecular and Cellular Biology, Jefferson 254, 17 Oxford Street, Cambridge, MA 02138 USA. Electronic mail: [email protected] A fabrication method for positioning and embedding a single-walled carbon nanotube (SWNT) across the diameter of a solid state nanopore is presented. Chemical vapor deposition (CVD) is used to grow SWNTs over arrays of focused ion beam (FIB) milled pores in a thin silicon nitride membrane. This typically yields at least one pore whose diameter is centrally crossed by a SWNT. The final diameter of the FIB pore is adjusted to create a nanopore of any desired diameter by atomic layer deposition (ALD), simultaneously embedding and insulating the SWNT everywhere but in the region that crosses the diameter of the final nanopore, where it remains pristine and bare. This nanotube-articulated nanopore is an important step towards the realization of a new type of detector for biomolecule sensing and electronic characterization, including DNA sequencing. 1 I. INTRODUCTION AND RATIONALE A nanopore through a solid state membrane or a biological cell membrane separating two ionic-solution filled compartments has shown itself to be a promising and versatile type of single molecule detector, capable of sensing the properties of both single small molecules and long polymers in solution.1-7 When a bias of 60 – 500 mV is applied across the membrane separating the two ionic solutions, charged polymeric molecules (e.g. DNA) in the solution are electrophoretically driven through the nanopore and each molecule can be detected as it traverses through the nanopore constriction. Until now, molecular detection has usually relied on observing a drop in the ionic conductivity as each molecule is driven through a single nanopore. The resolution and sensitivity of this molecule detector have been limited by the very small changes in ionic current and the high speed at which a DNA polymer translocates through the nanopore (usually >1 nucleotide/μsec for solid state nanopores).8,9 One is therefore led to consider other molecule detection modes that are capable of discerning structure along the molecule at higher resolution and greater electronic sensitivity. Two possibilities are to sense molecules with a nanoscale field effect transistor (FET) embedded across the pore diameter or by observing tunneling or other forms of electronic transport between gapped electrodes embedded in the nanopore.6,10 Important steps to fabricating gapped metallic electrodes atop a nanopore have recently been demonstrated.11,12 But instead of metallic probes, an electrically contacted carbon nanotube across the diameter of the nanopore is particularly attractive because of its chemical stability, nanoscale dimensions and highly ordered structure. Such a nanotube can serve as a 2 FET13,14 or, when suitably gapped, as the probes of a tunneling device.6,15-17 During molecular transport through such a device, electronic rather than ionic currents can be sensed, potentially offering improved signal to noise and a more controlled environment for molecular transport. But producing a nanopore with an embedded single walled carbon nanotube across its diameter is a significant challenge that has not previously been confronted. Controlled positioning at specific device locations remains one of the most significant hurdles in the application of carbon nanotubes18,19, and our goal was to devise a straightforward method that could easily, but reproducibly, produce many devices having an embedded single walled carbon nanotube across at least one nanopore of any chosen diameter. Here we demonstrate one way such devices can be realized. II. EXPERIMENTAL Figure 1 presents a schematic of the fabrication method. First, a row of 1μm x 1μm iron (Fe) catalyst pads for CVD growth of carbon nanotubes were patterned on a low stress silicon nitride (SiNx) membrane by electron beam lithography (EBL), electron beam evaporation, and metal lift-off. Then, rows of ~30nm diameter pores were milled through the membrane, usually in 11 x 11 arrays in the proximity of each of the 1μm x 1μm Fe pads with an FIB (Fig. 2a). The pores in each array were spaced 100 nm x 100 nm center-to-center, such that each array of 121 pores measured ~1μm x ~1μm edge-to-edge. Carbon nanotubes were then grown from the Fe pads by CVD.20 Typically at least one of the carbon nanotubes growing from the Fe pad reached the pore array and crossed the center of at least one of the FIB pores. Because the Fe catalyst pad 3 seeded the growth of multiple nanotubes that grew in many directions, we often noted several nanotubes that grew across other pores in the pore array. To adjust the nanopore size and embed the nanotube in the finished nanopore, the samples were transferred to an ALD chamber that was used to conformally deposit multiple layers of insulating Al2O3 on all exposed surfaces, including the perimeter of the nanopores, but not on the portion of the nanotube that is to serve as an FET sensor or tunneling probe. ALD of high- dielectric materials has been shown to produce a benign dielectric/SWNT interface that does not adversely affect the electrical properties of the nanotube.21,22 Because ALD of Al2O3 is a self-limiting process that depends on half- cycle reactions that proceed through the formation of Al–O Lewis acid-base complexes,23 the hydrophobic surface of a pristine defect-free carbon nanotube, in which all the carbon atoms are in sp2 configuration, does not initiate the growth of aluminum oxide from its surface.22,24,25 Rather, the Al2O3 layers grew from all the silicon nitride surfaces, with lateral growth from the newly formed Al2O3 eventually burying only those portions of the carbon nanotube lying on the silicon nitride surface and near the inside diameter of the FIB pore. Deposition of a known number of Al2O3 layers, each of which is ~0.1 nm thick, on the inside diameter of the FIB pore shrank the pore diameter to a predictable smaller size,26 with the length of exposed nanotube shrinking commensurately. Only the portion of the suspended carbon nanotube that passed across the finished ALD coated nanopore remained uncoated (Figure 3). Thus the Al2O3 coating could serve as a protective layer that insulates the nanotube from conducting fluid and the influence of proximal molecules anywhere in solution except in the finished nanopore. 4 We found the following procedural details consistently produced 2–10 nm nanopores whose diameters were crossed by one single-walled carbon nanotube. We started with 80 x 80 μm free-standing, low-stress 250-270 nm thick SiNx membranes, obtained by anisotropic KOH etching of a SiNx coated silicon wafer.27 To assure cleanliness, the silicon chips bearing the membranes were cleaned by dipping in trichloroethylene, acetone, and finally methanol, and baked at 150oC in air. Subsequently, a row of 0.5 to 1 nm thick 1μm x 1μm Fe pads were patterned on the membrane using standard EBL processing (Raith-150 system) with a Poly(methyl methacrylate) (PMMA) resist that was lifted off with acetone followed by a rinse in isopropyl alcohol and drying in a N2 atmosphere. The pore arrays were milled using a FEI Micrion 9500 FIB system operated at 50 kV acceleration voltage, with an aperture diameter of 15 m corresponding to a 1.4 pA beam current. To obtain the smallest possible diameters of the pores, a single point (1 pixel) ion beam was rastered through the array ~100 times to obtain a total beam time of 6s/pore. The resulting pores typically had ~35 – 40 nm diameters and were spaced 100 nm apart in a square array. Each array, with typically 11 x 11 pores, was located 1 μm away from the pre-made Fe pads. Figure 2(a), shows a transmission electron microscope (TEM) image of an array of FIB milled pores (JEOL 2100 operated at 200 KV). In this example, the pores have a diameter of ~40 nm and it can be seen that the centered 11 pore x 11 pore array is but one of several similar pore arrays which were patterned in a row parallel to the row of pre-made Fe catalyst pads (not shown). It is noted that the area of the membrane between the pores is partially milled because the ion beam is not blanked as it moves from one point to the next. 5 Carbon nanotubes were grown from the Fe pads in a 1 inch diameter quartz tube furnace (Lindberg/Blue Mini-Mite 1100 C), with methane gas (CH4) as the carbon source. First, the sample temperature was ramped to 900oC in a flow of 500 sccm of Ar. When it reached 900oC, the sample was annealed in 200 sccm of pure H2 for 10 mins. Carbon nanotube growth was initiated by introducing a 1000 sccm flow of CH4 for 15 minutes together with the 200 sccm H2 flow. After growth, the system was cooled down to room temperature in a 500 sccm flow of Ar. These conditions usually yielded the growth of single-walled carbon nanotubes that were several microns in length with a high probability of at least one nanotube from each Fe catalyst pad growing across a nanopore array. Through many trials, we found the distances between the Fe-pads and the pore arrays and the spacing between the pores in each array stated above yielded a high probability that a nanotube would centrally cross the diameter of at least one pore (Figure 2b). Atomic layer deposition of Al2O3 was conducted in a home-built system at 225oC, by alternating cycles of flowing trimethylaluminum (TMA) vapor followed by water vapor. At the end of each cycle, one atomic layer of Al2O3 (thickness ~0.1 nm) was deposited.5 As shown in Figure 3b (arrows), the interface between the initial pore perimeter and the deposited Al2O3 layer can be seen in electron micrographs and it is clear that the carbon nanotube over the finished nanopore remained uncoated after the ALD had shrunk the pore to the finished desired diameter. In this example, the final nanopore diameter was 11 nm, the added alumina thickness was 10 nm, and the nanotube diameter was 2.6 nm, but by using fewer or more ALD cycles, the same procedure was used to produce many other larger or smaller diameter nanopores crossed by an 6 embedded nanotube. To keep the nanotube across the finished nanopore free from alumina deposition, direct exposure to an electron beam during SEM or TEM observation (as was done to produce the image in Figure 2b) must be avoided before the ALD processing. This is because direct exposure to an e-beam inevitably contaminates or causes defects in the nanotube which then nucleate alumina deposition. With appropriate electrical connections to the carbon nanotube, the exposed length of a semiconducting nanotube that crosses the finished nanopore diameter (Figure 3b) can serve as an FET single molecule detector because the translocation of a molecule in its immediate vicinity will influence the carrier concentrations in the nanotube. Alternatively, a gapped nanotube with ends abutting the perimeter of the finished nanopore can serve as a tunneling device to sense and characterize molecules as they pass through the nanopore. To achieve the latter configuration, the uncovered portion of the nanotube crossing the nanopore (Figure 3b) can easily be eliminated by exposure to a low energy (e.g. 0.5-1 keV) sputter ion beam or plasma etch apparatus, which will rapidly remove the exposed length of the nanotube while leaving the alumina-covered portion of the original nanotube untouched. For illustrative purposes, even the electron beam of the TEM itself can be used, as can be seen in Figure 3c. But a low energy ion beam is preferred because its very short penetration distance into the alumina will minimize damage to the alumina covered portion of the gapped nanotube. Other methods of removing the suspended part of the nanotube over the nanopore, such as reactive ion etching, are currently being investigated. 7 III. DISCUSSION AND CONCLUSION Using an ungapped or gapped carbon nanotube integrated with a nanopore will require electrically connecting the appropriate positions of a selected nanotube to the outside world before they are buried and insulated by ALD. This can be accomplished without device damage or contamination by ice lithography28 before the ALD procedure. As noted above, we have found that carbon nanotube exposure to an electron beam during standard EBL and lift-off damages or contaminates the nanotube. An unwanted consequence of such contamination or damage is that subsequent ALD coats the entire length of the nanotube,25 including the portion that must remain uncoated if it is to serve as an FET or be gapped by an ion beam in the finished nanopore. Ice lithography avoids this problem because imaging, locating, and selecting the desired nanotube and determining the desired pattern of metal contacts are accomplished while the entire device, including the nanotube, is protected under a thin layer of water ice.29 Furthermore, we and others have found that ALD of Al2O3 and other high-dielectric materials produces a benign dielectric/SWNT interface that does not adversely affect the electrical properties of the nanotube.22,30 Finally, we note that if the electrical signals from only the contacted nanotube that successfully crosses the diameter of a single nanopore is used to sense the presence of molecules translocating through that nanopore, the presence of unused nanopores in the pore array is of no concern. In summary, we have demonstrated a simple and reproducible nanoscale fabrication technique for articulating a nanopore with a carbon nanotube. Although producing arrays containing a large number of such articulate nanopores cannot depend on individually selected e-beam lithographic processing and will require more elaborate, 8 controlled methods to position nanotubes across the diameter of multiple nanopores, the simple approach we have demonstrated here is an important step towards the realization of a nanodevice platform whose value for in-solution sensing and characterization of molecules and biopolymers needs to be established before more elaborate, highly parallel and better controlled fabrication methods are implemented. ACKNOWLEDGMENTS This work was supported by NIH award number R01HG003703 to J. Golovchenko and D. Branton. REFERENCES 1J.J. Kasianowicz, E. Brandin, D. Branton, and D.W. Deamer, Proc. Natl. Acad. Sci. U.S.A. 93, 13770 (1996). 2J. Li, D. Stein, C. McMullan, D. Branton, M.J. Aziz, and J.A. Golovchenko, Nature 412, 166 (2001). 3J. Li, M. Gershow, D. Stein, E. Brandin, and J. Golovchenko, Nature Materials 2, 611 (2003). 4A.J. Storm, J.H. Chen, X.S. Ling, H.W. Zandbergen, and C. Dekker, Nature Materials 2, 537 (2003). 5P. Chen, J. Gu, E. Brandin, Y.-R. Kim, Q. Wang, and D. Branton, Nano Lett. 4, 2293 (2004). 9 6D. Branton, D. Deamer, A. Marziali, H. Bayley, S.A. Benner, T.Z. Butler, M. Di Ventra, S. Garaj, A. Hibbs, X. Huang, S.B. Jovanovich, P.S. Krstic, S. Lindsay, X.S. Ling, C.H. Mastrangelo, A. Meller, J.S. Oliver, Y.V. Pershin, J.M. Ramsey, R. Riehn, G.V. Soni, V. Tabard-Cossa, M. Wanunu, M. Wiggin, and J.A. Schloss, Nature Biotechnology 26, 1146 (2008). 7S. Garaj, W. Hubbard, A. Reina, J. Kong, D. Branton, and J.A. Golovchenko, Nature 467, 190 (2010). 8D. Fologea, J. Uplinger, B. Thomas, D.S. McNabb, and J. Li, Nano Lett. 5, 1734 (2005). 9D. Fologea, M. Gershow, B. Ledden, D.S. McNabb, J.A. Golovchenko, and J. Li, Nano Lett. 5, 1905 (2005). 10M. Zwolak and M. Di Ventra, Reviews of Modern Physics 80, 141 (2008). 11M. Taniguchi, M. Tsutsui, K. Yokota, and T. Kawai, App. Phys. Lett. 95, 123701 (2009). 12A.P. Ivanov, E. Instuli, C.M. McGilvery, G. Baldwin, D.W. McComb, T. Albrecht, and J.B. Edel, Nano Lett. 11, 279 (2011). 13G. Gruner, Anal. Bioanal. Chem. 384, 322 (2006). 14B.L. Allen, P.D. Kichambare, and A. Star, Adv. Mater. 19, 1439 (2007). 15M. Zwolak and M. Di Ventra, Nano Letters 5, 421 (2005). 16H. Tanaka and T. Kawai, Nature Nanotechnology 4, 518 (2009). 17M. Tsutsui, M. Taniguchi, K. Yokota, and T. Kawai, Nature Nanotechnology 5, 286 (2010). 10 18S. Banerjee, B.E. White, L. Huang, B.J. Rego, S. O'Brien, and I.P. Herman, J. Vac. Sci. Technol. B 24, 3173 (2006). 19M. Duchamp, K. Lee, B. Dwir, J.W. Seo, E. Kapon, L. Forro, and A. Magrez, ACS Nano 4, 279 (2010). 20H.B. Peng, T.G. Ristroph, G.M. Schurmann, G.M. King, J. Yoon, V. Narayanamurti, and J.A. Golovchenko, Applied Physics Letters 83, 4238 (2003). 21A. Javey, J. Guo, D.B. Farmer, Q. Wang, E. Yenilmez, R.G. Gordon, M. Lundstrom, and H.J. Dai, Nano Lett. 4, 1319 (2004). 22D.B. Farmer and R.G. Gordon, Nano Lett. 6, 699 (2006). 23Y. Widjaja and C.B. Musgrave, App. Phys. Lett. 80, 3304 (2002). 24D.B. Farmer and R.G. Gordon, Electrochemical and Solid-State Letters 8, G89 (2005). 25D.B. Farmer, Chapter 6 of PhD thesis, Harvard University (2007). 26P. Chen, T. Mitsui, D.B. Farmer, J. Golovchenko, R.G. Gordon, and D. Branton, Nano Lett. 4, 1333 (2004). 27K.E. Bean, IEEE Transactions on Electron Devices 25, 1185 (1978). 28G.M. King, G. Schurmann, D. Branton, and J.A. Golovchenko, Nano Letters 5, 1157 (2005). 29A. Han, D. Vlassarev, J. Wang, J.A. Golovchenko, and D. Branton, Nano Lett. 10, 5066 (2010). 30A. Javey, J. Guo, D.B. Farmer, Q. Wang, D. Wang, R.G. Gordon, M. Lundstrom, and H. Dai, Nano Letters 4, 447 (2004). 11 Figure Captions Figure 1. (Color online) Schematic diagram of the fabrication process of a carbon nanotube articulated nanopore. (a) Deposition of iron (Fe) pads on a silicon nitride membrane by a combination of electron beam lithography patterning and metal lift-off. (b) Arrays of nanopores are milled through the SiNx membrane by a FIB machine. (c) Chemical vapor deposition growth of a carbon nanotube from the Fe pad over the array of pores. (d) Zoom-in view of a nanopore from the array with a carbon nanotube crossing its diameter. (e) A layer of aluminum oxide (Al2O3) is deposited by ALD. (f) Zoom-in view of the carbon nanotube articulated nanopore after ALD showing the decrease of its diameter by ALD growth of Al2O3 from the pores initial perimeter. Figure 2. Transmission electron microscopy images of (a) an FIB milled array of pores drilled through a SiNx membrane. (b) A FIB milled pore from an array after a single- walled carbon nanotube has grown across the pore’s diameter. Scale bar is 10 nm. Figure 3. Transmission electron microscopy images. (a) A nanopore array after the deposition of Al2O3 by ALD. The circular pattern of dots here and in b and c denotes the boundary of the initial FIB pore perimeter which encloses the subsequently deposited Al2O3 layers. (b) TEM image of a nanotube articulated nanopore after ALD deposition. (c) TEM image of the same nanopore as in b, after the portion of the nanotube not covered with Al2O3 has been ablated with the microscope’s electron beam. 12 Figure 1 13 Figure 2 14 Figure 3 15
1904.01040
1
1904
2019-04-01T18:06:41
Bayesian inference for nanopore data analysis
[ "physics.bio-ph", "physics.data-an" ]
Nanopore sensors detect the substructure of individual molecules from modulations in an ion current as molecules pass through them. In this work, we present the classification of features in the substructure as a case study to illustrate the power of Bayesian inference when analysing nanopore data. A brief introductory section provides an overview of the core concepts, followed by a detailed description of the analysis procedure to facilitate other researchers to add Bayesian inference to their toolbox. Our hybrid approach of a classical peak-finding algorithm and Bayesian model comparison allows the probabilistic classification of features as "0" or "1" bits by calculating relative evidences for two competing models. We correctly classify on average ~ 70% of bits for individual events and use the probabilistic nature of the approach to calculate a cumulative estimate with an accuracy of > 94%. The technique presented here is readily extensible to models of the translocation process which can take into account arbitrary molecular designs, our approach may therefore be used to analyse a wide range of features observed in nanopore experiments.
physics.bio-ph
physics
Bayesian inference for nanopore data analysis Niklas Ermann, Kaikai Chen, and Ulrich F. Keysera) Cavendish Laboratory, University of Cambridge, 19 JJ Thomson Avenue, Cambridge CB3 0HE, UK Nanopore sensors detect the substructure of individual molecules from modulations in an ion current as molecules pass through them. In this work, we present the classification of features in the substructure as a case study to illustrate the power of Bayesian inference when analysing nanopore data. A brief introductory section provides an overview of the core concepts, followed by a detailed description of the analysis procedure to facilitate other researchers to add Bayesian inference to their toolbox. Our hybrid approach of a classical peak-finding algorithm and Bayesian model comparison allows the probabilistic classification of features as "0" or "1" bits by calculating relative evidences for two competing models. We correctly classify on average ∼ 70% of bits for individual events and use the probabilistic nature of the approach to calculate a cumulative estimate with an accuracy of > 94%. The technique presented here is readily extensible to models of the translocation process which can take into account arbitrary molecular designs, our approach may therefore be used to analyse a wide range of features observed in nanopore experiments. I. INTRODUCTION A Resistive-pulse sensing based on nanopores is utilised not only in the detection but increasingly in the study of substructure of single molecules1,2. Characterisation of molecular substructure relies on accurately identifying features within a translocation signal. For example, such features in the ionic current could correspond to DNA folds3, knots4 or modifications such as protrusions5,6 or attached proteins7. Modifications offer the possibility to store digital data8 or indicate the presence of certain base sequences9. In the example shown in Figure 1, each modification produces a secondary current drop in the already reduced current, creating a bit sequence of high "1" and low "0" signals which can be used to encode in- formation. In the ideal scenario, the amplitude of a drop would relate exactly to the bit encoded at this specific position. In a real nanopore system, however, several sources of noise complicate the decoding process10. The bottom part of Figure 1 illustrates that fluctuations in the velocity of a translocating DNA strand11 shift the position of secondary drops, preventing the exact local- isation of bits purely based on their appearance in the signal. Variations in the drop amplitude create ambigu- ity as to which modification produced the feature, which is exacerbated by the Gaussian noise inherent in the cur- rent measurement10. As a result, analysis procedures which aim to extract the molecular substructure that gives rise to the features within a translocation signal need to be able to distinguish between signal and noise and quantify the probability that a certain structure has been detected. Existing approaches to nanopore data analysis have been dominated by classical algorithms. Several tech- niques have focused on finding the true magnitude and duration of the current drop when constrained by the finite filter rise times inherent in instrumentation12,13. a)Electronic mail: [email protected] t u o d a e r l a e d I i e s o N t n e r r u c time 010000110100000101010110 "C" "A" "V" Gaussian noise Velocity fluctuations Amplitude variations FIG. 1. Experimental noise complicates the readout of bit sequences stored on a double-stranded DNA scaffold. In the ideal case, bits can be assigned directly from the amplitude of the secondary current drops produced by modifications. The upper idealised current trace shows that the letters "CAV" can be easily decoded in 8-bit ASCII. The simulated bot- tom trace demonstrates that Gaussian noise inherent in the measurement, velocity fluctuations of the DNA strand and variations in peak amplitudes complicate the decoding. Raillon et al. developed a cumulative sums algorithm to identify sequential steps during a translocation14. Fur- ther efforts went into the detection of secondary current drops as well as the development of comprehensive soft- ware tools for the analysis of nanopore data6,15,16. While these classical techniques can provide fast and reliable event characterisation for their specific use case, their performance often relies on user-defined inputs and ar- bitrary thresholds. Recent work on convolutional neural networks has shown that they successfully classify events in a generalised fashion, at superior data usage and ac- curacy compared to conventional techniques17. However, 9 1 0 2 r p A 1 ] h p - o i b . s c i s y h p [ 1 v 0 4 0 1 0 . 4 0 9 1 : v i X r a the approach requires labelled datasets containing thou- sands of events for each molecular species to be detected. Bayesian inference is a promising alternative for the extraction of information from experimental data18. In the past, its use has been hampered by the computa- tional cost associated with calculating high-dimensional integrals. This hurdle has been partially overcome by advances in computational power over the last decades as well as the development of efficient Markov chain Monte Carlo-based (MCMC) algorithms19. The prob- abilistic nature of Bayesian techniques allows the calcu- lation of probability distributions for model parameters, as opposed to the single-value estimates often found in classical approaches. The probability landscapes directly show a wealth of features which may otherwise remain obscured, such as multimodal distributions and regions surrounding maximum-likelihood parameter estimates20. In addition to the characterisation of model parameters, the Bayesian approach assigns a probability to the valid- ity of the model itself. Called evidence, this value allows comparing different candidate models to assess which one best explains the experimental data and thus which phys- ical theory should be favoured21. In this paper, we demonstrate that the probabilistic nature of Bayesian inference allows us to draw unbi- ased conclusions about the substructure of translocating molecules from noisy current data. As a case study, we use recently published data from a double-strand of DNA which has been modified with differently sized DNA over- hangs at defined positions along its length8. As shown in Figure 1, each modification produces a secondary current drop in addition to the already reduced current whose magnitude depends on the length of the overhang. The total number of overhangs is defined through the design of the molecule, however the bit sequence remains un- known. The question for data readout thus becomes: given a number of events, what is the relative probabil- ity of an overhang at a certain position being long or short, i.e. signifying a "0" or "1" bit? In the follow- ing we show that a Bayesian approach is ideally suited to analyse substructure in nanopore translocation data. Before illustrating the practical procedure, we give a brief introduction into Bayesian inference and reasoning. II. BAYESIAN INFERENCE Bayesian techniques emerge from a fundamental state- ment about conditional probabilities called Bayes' rule22: likelihood prior z p(Dθ, M ) p(θM ) z } { p(DM ) } { {z } evidence Here θ denotes the model parameters, D the experi- mental data and M a certain model. Bayes' rule tells p(θD, M ) = posterior {z } (1) 2 us that the probability for the model parameters to take on a certain value given the data, called the posterior, is proportional to the probability of obtaining the data from the given choice of parameters, called the likelihood. While the two probabilities sound similar, this is in fact a powerful statement: in data analysis we are interested in the posterior, i.e. we want to know what the data can tell us about the parameters of our explanatory model. Bayes' rule allows us to obtain the posterior by relating it to a probability we can calculate, the likelihood. Us- ing a nanopore signal as an example, Figure 1 shows that the current trace exhibits Gaussian noise whose variance we call σ2 n. We represent any model M describing the data with a function h(θ) which defines how the model relates its parameters θ to the observed signal. The like- lihood for an individual data point d given θ and M is then simply proportional to the Gaussian noise probabil- ity distribution centered around d at the point predicted by h(θ), that is p(dθ, M ) ∝ exp(cid:18) − (h(θ) − d)2 n (cid:19) 2 · σ2 (2) Having obtained an expression for the likelihood, the remaining term with a dependence on the parameters θ is the prior p(θM ). It encodes previous knowledge about the parameters, for example from the posteriors obtained from previous experimental data20. A lack of initial infor- mation about θ is easily represented by a flat probability distribution. The prior in combination with the likeli- hood lets us calculate the quantity of interest, namely the posterior probability over the parameter values given the data. An accessible treatment of this approach can be found in20, including examples on how the Bayesian view motivates widespread statistical techniques such as least squares regression. Up to this point, the discussion has focused on deter- mining the parameter probability distribution for a given model. However, within the probabilistic framework we can equally ask which of several models is most likely given a set of experimental data. To answer this ques- tion we need to calculate p(MiD) where i indexes several competing models. Bayes' rule tells us that p(M D) = p(DM )p(M ) p(D) ∝ p(DM ) In the last term of the equation we recognise the evi- dence p(DM ) which appears in Equation 1. Assuming uniform priors over different models Mi, the model with the highest probability of explaining the data is thus the one with the largest evidence. To compute the evidence, we note that a) Bit sequence 00000000111111110100010011111111011011110000000011111111 3 Short overhang: '0' bit DNA scaffold Long overhang: '1' bit (3) (4) dθ p(DM ) =Z p(D, θM )dθ =Z p(Dθ, M ) } {z likelihood p(θM ) prior {z } The evidence is therefore an integral of likelihood and prior over the entire parameter space. Particularly for high-dimensional models where analytical optimisations aren't possible, it is these integrals which make Bayesian techniques computationally intensive. However, once the calculation becomes feasible we obtain a probabilistic method to assess the explanatory power of different mod- els. In the next section we show how this Bayesian tool allows us to extract unbiased information from nanopore data. The interested reader is referred to18 and21 for more detailed discussions of Bayesian inference. b) ) A n ( t n e r r u C 1 . 0 - 3 . 0 - III. CASE STUDY: BIT SEQUENCE IDENTIFICATION FROM NANOPORE DATA As mentioned in the introduction, we use previously published data from a modified double-strand of DNA to illustrate how nanopore data analysis can benefit from Bayesian techniques8. Each of the overhangs along the strand's length produces a secondary current drop whose amplitude depends on the overhang's size (Figure 2). In total the strand has been decorated with 56 such modi- fications, resulting in a sequence of 56 bits. For each bit, the goal is to estimate whether it corresponds to a "0" or "1", depending on the size of the overhang. The exper- imental data was obtained from a homogeneous sample with each strand modified in the same way. This means that we can sequentially build up the estimate for each bit as more events are added to the total dataset. We denote the data for a single event i by Di and the to- tal dataset encompassing all events by {Di}. Each Di consists of current data y and time points x. We initially carry out several preprocessing steps. These use classical algorithms and return the intra-event times and amplitudes of the 56 most significant secondary current drops. We discuss the procedure in the following to present the complete analysis pipeline. A. Preprocessing steps The goal of the initial data processing is to provide values for the priors used in the comparison of the "0" bit and "1" bit models in the next section. First, we de- termine current levels corresponding to the unimpeded pore L0 and one double strand in the pore L1. This is achieved by compiling the concatenated current data from all events into a kernel density estimate and detect- ing the two peaks in the distribution as described in23. We define a cut-off value C as C = L0 − 0.75 ∗ (L0 − L1) 0.0 x1 * * ... x2 1.0 Time (ms) 2.0 * x56 FIG. 2. Modifications along a double-stranded DNA scaffold encode 56 bits, as published in8. (a) Schematic of the DNA molecule. Short DNA overhangs along the scaffold's length encode "0" bits (blue), long overhangs represent "1" (orange). (b) An example current trace produced by the molecule in (a). Red dots indicate the secondary current drops selected by a classical peak-finding algorithm, each drop corresponds to one of the 56 overhangs. We determine the drop ampli- tudes relative to the shaded green region, which is calculated from an auxiliary trace where the drops have been removed (continuous green line in the inset). which will be used later to define the start and end of each event. Next, we carry out the following steps for each event Di with current trace y and time points x: 1. Secondary current drop identification We detect the secondary current drops for each event using a peak-finding algorithm adapted from15. We are tolerant of false positives at this point and may obtain more than 56 drops. These will be filtered out in a later step. We discard events for which the peak finding results in fewer than 56 drops. 2. Identification of intra-event baseline current. We use programmatic notation in the following, where ¯v[i : j] refers to the slice of vector ¯v between indices i and j. Assignments of the form A = B work from right to left, i.e. assign the value B to the variable A. (a) We first remove the secondary current drops identified in the previous step by setting the current values in each subregion found to be part of a current drop to the most positive value within that subregion. This gives us a new current trace y with the drop section re- placed by piecewise constant values as shown by the green line in the inset in Figure 2 b). (b) On the modified current trace y we determine two index values a and b which define the first and last positions at which y < C, where C is the cut-off value calculated initially. We then fit a linear function f (x, s, i) = i + s ∗ x to the region y[a : b], obtaining least-squares fit values i and s. (c) On the original current trace y, we again de- termine two index values a and b correspond- ing to the first and last positions at which y < C. We then replace the slice between these indices with values calculated from the function f (x, s,i) from the previous step, i.e. if the new current trace y is initially the same as y, we assign y[a : b] = f (x[a : b],i, s) to obtain a new current trace y that contains a linear intra-event baseline between the start and end of the event without taking into ac- count secondary drops. This trace, shown by the shaded green region in Figure 2 b), serves as the baseline from which we calculate the amplitudes of the secondary drops. 3. Drop filtering In the case of more than 56 drops, we keep the 56 with the largest distance from the baseline y. In the example event in Figure 2 b) the selected drops are marked with red dots. This simple filtering step makes the system susceptible to shift errors, which occur when spurious drops shift the position as- signments of all other bits. Such errors as well as techniques to avoid them are discussed in section III B. Having identified drops for each event, we calculate ag- gregate values across the whole dataset {Di} excluding events for which fewer than 56 drops could be detected. We distinguish two populations in the drop amplitude corresponding to "0" and "1" bits. By fitting Gaussians to the peaks in the kernel density estimate of all drop amplitudes within an event we obtain estimates of the mean secondary drop amplitude mj and the spread in amplitudes wj where j = 0, 1 for "0" and "1" bits re- spectively. It should be noted that the two populations can only be clearly distinguished for some events, we use the arithmetic mean for all events for which identification is possible to obtain the average mean and width of the amplitude drops for each nanopore, ¯mj and ¯wj. 4 B. Bayesian model comparison The above preprocessing steps provide us with the positions of the most prominent 56 current drops for each event, as well as the distribution of drop ampli- tudes across the whole dataset. We use a classical algo- rithm to identify current drops, as opposed to running Bayesian inference on a large model that fully incorpo- rates all features of a translocation event. The classifi- cation of 56 bits means such a model would contain at least 56 parameters. State-of-the-art nested sampling al- gorithms that allow full Bayesian inference scale at worst as E ∼ O(P 3) or even exponentially, where E is the number of likelihood evaluations and P is the number of model parameters24. A large number of parameters thus strongly increases the computation time, which makes the full model approach unfeasible for a higher number of bits. The computational cost is exacerbated by the vari- ation in the times between consecutive drops due to fluc- tuations in the translocation velocity of the DNA strand. Designing "1" and "0" bits as differently sized overhangs as opposed to the presence and absence of modifications somewhat mitigates this issue and leads to a greatly im- proved ability to identify current drops. However, at- tempts to create a full model showed that a large num- ber of additional parameters is still required to correctly detect drops when the distance between them cannot be constrained to one value. For the above reasons we follow a hybrid approach by using a classical algorithm to identify the 56 most prominent current drops and Bayesian inference to con- secutively assign a bit to each drop, which allows us to take advantage of Bayesian probabilities without exces- sive computational demands. To do so, we use the in- formation obtained in the preprocessing steps to define two models, M0 and M1, corresponding to a single "0" and "1" bit respectively. Within each event, we then use the Bayesian approach to compare the models' evidence values, thereby classifying each current drop as "0" or "1". From Equation 3 it is apparent that we require an ex- pression for the likelihood to calculate evidences. As- suming Gaussian noise in the current signal, we define the likelihood as P (Dθ, Mj) ∝ exp(cid:18)− L Xk=1 (h(xk, yk, θ) − yk)2 2 · σ2 n (cid:19) where xk, yk and yk are the kth elements of the time and the initial and corrected current data respectively and the second line is directly comparable to Equation 2. The sum over the trace elements stems from the assump- tion of independent samples, meaning we can multiply the Gaussian noise distributions for each element. Within the exponent, L is the length of the current trace and σn the Gaussian noise in the current. The function h(x, b, θ) fully defines how the model Mj relates a) ) A n ( t n e r r u C b) ) R ( n l 0.1  0.3 0.5  20 0 20  5 '0' bit '1' bit ln(R) > 0: '1' bit ln(R) < 0: '0' bit 1.5 0.5 1.0 Time (ms) FIG. 3. Bayesian model comparison assigns bits to each of the 56 current drops in the single event shown. (a) Illustration of the analysis procedure on the translocation event from Figure 2 (b). Stepping through the identified current drops, we compare the evidences for two models representing a "0" and "1" bit. The orange traces show the model outcomes for maximum likelihood parameter estimates on two exemplary drops, continuous and dashed lines correspond to the "0" and "1" bit models respectively. It is clear that the "1" model more accurately describes the data for the first drop, whereas the "0" model fits the second drop. (b) Outcome of the Bayesian model comparison. ln(R) describes the logarithm of the ratio of the evidences for the "1" and "0" models, ln(R) > 0 indicates a "1" bit and vice versa. The analysis correctly classifies 49 out of 56 bits in the event shown, red bars and arrows show wrong assignments. its parameters θ to measured data. We represent the sec- ondary current drops by Gaussian functions with mean µ, amplitude a and width σ added to a baseline b, i.e. h(x, b, θ) = h(x, b, µ, a, σ) = a · exp(cid:18) − (x − µ)2 2 · σ2 (cid:19) + b In the analysis presented here, this model is used for both "0" and "1" bits, with the distinction between the two coming from different priors on the amplitude pa- rameter a. This is an empirical simplification as a de- tailed physical understanding of how modifications pro- duce their associated current drops is the subject of on- going research. In addition, at this point it is not known to what extent the differently sized overhangs influence other drop characteristics such as shape. As more is un- derstood about the underlying process and the experi- mental data the two models should be refined. In addition to the likelihood, we require prior distribu- tions over the model parameters θ. We choose uniform, normalised priors, as described in the appendix to24. We l −0.5·∆x∗, x∗ constrain the values for the mean µ for the lth current drop to the interval [x∗ l +0.5·∆x∗], where x∗ l is the time position of the lth drop and ∆x∗ is the mean difference between all drop positions within the event (see Figure 2). A prior in the interval [0.002 ms, 0.006 ms] ac- curately describes the width σ of the secondary current drops. As mentioned above, the distinction between the two hypotheses comes from different priors on the ampli- tude a of the Gaussian function: for aj,min < a < aj,max otherwise. P (aMj) =((aj,max − aj,min)−1 0 where aj,min = ¯mj + sj,min · ¯wj aj,max = ¯mj + sj,max · ¯wj and ¯mj and ¯wj are the mean and width of the cur- rent drop populations identified at the end of section III A. The factors in the prior limits were chosen as s0,min = −0.5, s0,max = +0.5, s1,min = −2 and s1,max = 0 to optimally represent the current drop amplitudes for the two bits j = 0, 1. A simple Gaussian fit to each cur- rent drop and classification according to an amplitude threshold would be less computationally intensive. How- ever, as mentioned above the approach presented here is generic in that it can easily be extended to take into account current drop characteristics such as shape by adapting the model function h(x, b, θ). Secondly, we ob- tain evidences for each model, which means the classifi- cation into "0" and "1" bits is easily combined into an aggregate value across many events. Having obtained the likelihood and prior, we use the MultiNest Bayesian inference algorithm to compute log evidence values ln(cid:0)P (DMj)(cid:1)25. The inset in Figure 3 a) illustrates our sequential approach to bit assignment for two exemplary bits: at the position of each bit, we calcu- late evidences for the competing models, select the clas- sification with the higher relative probability and move on the next position. The orange traces show the model outcomes using maximum likelihood parameter estimates which we obtain as part of the evidence calculation. It is clear that the "1" model matches the first current drop more accurately, whereas the "0" model is preferred for the second drop. The Bayesian approach represents the concept of a 'better match' by the ratio of the posterior probabilities over the two models, R, which is related to the log evidences via ln(cid:0)R(cid:1) = ln(cid:18) P (M1D) P (M0D)(cid:19) = ln(cid:18) P (DM1)P (M1) P (DM0)P (M0)(cid:19) P (DM0)(cid:19) = ln(cid:18) P (DM1) = ln(cid:0)P (DM1)(cid:1) − ln(cid:0)P (DM0)(cid:1) where in the second line we assume equal priors for the two models, i.e. P (M1) = P (M0). For each current drop, we therefore calculate the difference between the log evidences for a "1" bit and a "0" bit to obtain the log of the probability ratio. The resulting value ln(R) pro- vides both a decision criterion between the two bits and an indication how strongly one hypothesis is favoured: ln(R) < (>)0 indicates a preference for the "0" ("1") bit, while the absolute value ln(R) represents the pref- erence strength. Figure 3 b) shows the outcome of the calculation for all bits in a single event. Our approach misclassifies 7 out of 56 bits for this particular example as shown by the red bars and arrows. Analysis of more events showed that shift errors are a common source of wrong bit assignments. These occur when the classification of bits is correct, but the erro- neous insertion or deletion of bits lead to the wrong as- signment of subsequent bits to positions in the sequence. Such errors are particularly problematic when the goal is to estimate the exact bit sequence for single events. If the 6 goal of the DNA modification is to create a library of dif- ferent sequences, a possible mitigation strategy is to not use all 56 available dimensions but create sequences with a maximal distance in vector space. For an estimated bit sequence shift errors can then be taken into account to find the most likely classification. A more direct ap- proach is to design overhangs so that they appear in bit "blocks" with detectable spacing between them, which limits the effect of erroneously identified current drops to the length of the block. Similarly, larger "beacon" overhangs allow verification of the position at certain in- tervals. An additional way to improve the classification accu- racy is to combine the bit estimates from multiple events. One of the advantages of the Bayesian approach is that it directly allows updating the bit estimate for each of the 56 positions as new events are added. As the DNA molecule can enter with either of its two ends first, we orient each event based on the number of "0" bits found in the 8 first and 8 last positions in the estimated bit sequence. It should be noted that apart from this correc- tion, our classification includes no prior information on the bit sequence. If {Di} is again the dataset of all events and N is the number of events, the overall evidence for hypothesis j is given by P(cid:16){Di}Mj(cid:17) = N Yi=1 P(cid:16)DiMj(cid:17) A derivation of this result can be found in section I of the supplementary information. The cumulative proba- bility ratio RN taking into account N events is then given by i=1 P (DiM1) ln(cid:0)RN(cid:1) = ln(cid:18) P (M1{D}) P (M0{D})(cid:19) = ln QN j=1 P (DjM0)! = ln N Yi=1 QN Xi=1(cid:16) ln(cid:0)P (DiM1)(cid:1) − ln(cid:0)P (DiM0)(cid:1)(cid:17) = P (DiM1) P (DiM0)! N where we again assumed equal priors for the two hy- potheses in the second line. This means that to obtain a cumulative log probability ratio for a number of events we simply add up their individual log probability ratios. The decision criterion remains the same as in the individ- ual case, where ln(RN ) < 0 indicates a preference for the "0" bit and vice versa, while the absolute value ln(RN ) represents how much one bit is favoured over the other. Figure 4 shows the aggregate log probability ratios for each bit position calculated from 1, 29 and 58 events. The inclusion of more experimental data decreases the error rate to 2 out of 56 bits, misclassifications are again N=1 ) 1 = N R ( n l 2 0 1 1 - 0 1 2 0 1 N=29 ) 9 2 = N R ( n l 2 0 1 1 - 0 1 2 0 1 N=58 ) 8 5 = N R ( n l 2 0 1 1 - 0 1 2 0 1 '1' '0' '1' '0' '1' 1 10 20 30 Bit # 4 '0' 50 FIG. 4. Combining bit estimates from multiple events im- proves the accuracy and confidence of assignments. ln(RN ) describes the logarithm of the evidence ratio for the two mod- els "0" and "1" compiled from N events, here N = 1, N = 29 and N = 58 from top to bottom. ln(RN ) > 0 again indi- cates a "1" bit and vice versa, the magnitude ln(RN ) shows how strongly one model is favoured. The number of wrong assignments (red bars) drops from 7 to 2 and the confidence increases as more event are included. shown by the red bars. The magnitude of the bars in- creases with N , indicating that the addition of further events leads to more confident estimates. It should be noted that wrong bit assignments depend on how many and which events have been included. The wrong or low- confidence estimates occur mainly in positions where a change occurs from "0" to "1" or vice versa. This can be explained by the shift errors outlined above: wrongly assigning current drops to shifted positions in the bit se- quence only produces an error if the neighbouring bit has a different value. The probability ratios obtained from the Bayesian approach thus directly indicate which bits are more difficult to classify, thereby pointing to the most likely source of errors and informing the design of improved DNA structures. To assess how the number of analysed events influ- ences the error rate, Figure 5 a) shows the assignment accuracy for individual events as well as for the estimate Cumulative Per event a y c a r u c c A 0 . 1 5 . 0 30 1 Analysed events N 60 b e c n e d i f n o c . m u c . m r o N 0 . 1 5 . 0 0 . 0 7 Overall Wrongly assigned 30 1 Analysed events N 60 FIG. 5. Analysing additional events improves the assignment accuracy as well as the confidence in bit estimates. (a) As- signment accuracy as a function of the number of analysed events N . While the individual (per event) accuracy fluctu- ates around 70%, the value for the cumulative estimate rises to above 90% within a few events. (b) Cumulative confi- dence as measured by the mean absolute log probability ratio h ln(RN )i. The blue line shows that the confidence including all bit estimates increases steadily as we analyse more events. The confidence calculated by including only wrong bit assign- ments, however, remains at a low level (red line). based on the cumulative log probability ratio ln(RN ). While the individual accuracy lies around 70% on aver- age (dashed orange line), the cumulative accuracy rises to above 90% within a few events and remains unaffected by consecutive low-accuracy estimates (blue line). Fig- ure 5 b) illustrates how the confidence in the assignments as measured by the mean absolute log probability ratio h ln(RN )i increases steadily as we analyse more events (blue line). Crucially, however, the confidence in wrong estimates does not rise substantially as more events are added (red line). This means that the increase in overall confidence is driven by more confident estimates mainly for correctly assigned bits, while estimates for wrongly classified bits remain uncertain. The Bayesian approach presented here thus correctly identifies large parts of the bit sequence while flagging wrong estimates with a low confidence. IV. CONCLUSION We have demonstrated that Bayesian inference is a powerful technique to analyse current data obtained from nanopore measurements. Using the readout of a bit se- quence encoded as modifications on a DNA strand as a case study, we show that our method correctly classifies > 94% of bits. Updating the probabilities as more events are taken into consideration follows naturally from the probabilistic nature of the Bayesian approach. The focus on probabilities further allows using evidence ratios as a confidence metric for each estimate. This provides valu- able information on which bits are difficult to assign and indicates the most likely sources of error. Bayesian infer- ence can therefore inform the design of improved DNA modifications, such as patterns to facilitate identification of the correct position in the bit sequence. While the Bayesian approach is a powerful tool for the readout of bit encodings, its generality makes it appli- cable to a wide range of analysis problems for nanopore data. In particular, one of its strengths lies in the com- putation of evidence values which probabilistically judge how well a model describes experimental data. As re- searchers continue to develop the physical understanding of the nanopore translocation process, competing theo- ries can easily be assessed through Bayesian model com- parison. In the context of DNA-carrier based nanopore sensing, this will shed light on open questions such as how the structure of modifications relates to the observed current drops. ACKNOWLEDGMENTS N.E. acknowledges funding from the EPSRC, Cam- bridge Trust and Trinity Hall, Cambridge. K.C. and U.F.K. acknowledge funding from an ERC consolidator grant (Designerpores 647144). REFERENCES 1P. Waduge, R. Hu, P. Bandarkar, H. Yamazaki, B. Cressiot, Q. Zhao, P. C. Whitford, and M. Wanunu, "Nanopore-Based Measurements of Protein Size, Fluctuations, and Conformational Changes," ACS Nano 11, 5706 -- 5716 (2017). 2A. Y. Y. Loh, C. H. Burgess, D. A. Tanase, G. Ferrari, M. A. McLachlan, A. E. G. Cass, and T. Albrecht, "Electric Single- Molecule Hybridization Detector for Short DNA Fragments," Analytical Chemistry 90, 14063 -- 14071 (2018). 3J. Li, M. Gershow, D. Stein, E. Brandin, and J. A. Golovchenko, "DNA molecules and configurations in a solid-state nanopore mi- croscope," Nature Materials 2, 611 -- 615 (2003). 4C. Plesa, D. Verschueren, S. Pud, J. van der Torre, J. W. Ruiten- berg, M. J. Witteveen, M. P. Jonsson, A. Y. Grosberg, Y. Rabin, and C. Dekker, "Direct observation of DNA knots using a solid- state nanopore," Nature Nanotechnology 11, 1 -- 6 (2016). 5C. Plesa, N. van Loo, P. Ketterer, H. Dietz, and C. Dekker, "Velocity of DNA during Translocation through a Solid-State Nanopore," Nano Letters 15, 732 -- 737 (2015). 6N. A. W. Bell and U. F. Keyser, "Digitally encoded DNA nanos- tructures for multiplexed, single-molecule protein sensing with nanopores," Nature Nanotechnology 11, 645 -- 651 (2016). 7C. Raillon, P. Cousin, F. Traversi, E. Garcia-Cordero, N. Hernandez, and A. Radenovic, "Nanopore Detection of Single Molecule RNAP-DNA Transcription Complex," Nano Letters 12, 1157 -- 1164 (2012). 8 8K. Chen, J. Kong, J. Zhu, N. Ermann, P. Predki, and U. F. Keyser, "Digital Data Storage Using DNA Nanostructures and Solid-State Nanopores," Nano Letters 19, 1210 -- 1215 (2019). 9R. M. M. Smeets, S. W. Kowalczyk, A. R. Hall, N. H. and C. Dekker, "Translocation of RecA-Coated Dekker, Double-Stranded DNA through Solid-State Nanopores," Nano Letters 9, 3089 -- 3095 (2009). 10R. M. M. and Proceedings of the National Academy of Sciences 105, 417 -- 421 (2008). Smeets, U. F. Keyser, N. H. Dekker, nanopores," C. Dekker, solid-state "Noise in 11A. J. Storm, C. Storm, J. Chen, H. Zandbergen, J.-F. Joanny, and C. Dekker, "Fast DNA Translocation through a Solid-State Nanopore," Nano Letters 5, 1193 -- 1197 (2005). 12D. Pedone, M. Firnkes, of ysis Analytical Chemistry 81, 9689 -- 9694 (2009). translocation events and U. Rant, "Data anal- experiments," in nanopore 13A. Balijepalli, J. Ettedgui, A. T. Cornio, J. W. F. Robertson, K. P. Cheung, J. J. Kasianowicz, and C. Vaz, "Quantifying Short-Lived Events in Multistate Ionic Current Measurements," ACS Nano 8, 1547 -- 1553 (2014). 14C. Raillon, P. Granjon, M. Graf, L. J. Steinbock, and A. Radenovic, "Fast and automatic processing of multi- level experiments." Nanoscale 4, 4916 -- 24 (2012). translocation nanopore events in 15C. Plesa and C. Dekker, "Data analysis methods for solid-state nanopores," Nanotechnology 26, 084003 (2015). 16J. H. Forstater, K. Briggs, J. W. F. Robertson, J. Ettedgui, O. Marie-Rose, C. Vaz, J. J. Kasianowicz, V. Tabard-Cossa, and A. Balijepalli, "MOSAIC: A Modular Single-Molecule Analysis Interface for Decoding Multistate Nanopore Data," Analytical Chemistry 88, 11900 -- 11907 (2016). 17K. Misiunas, N. Ermann, and U. F. Keyser, "QuipuNet: Convo- lutional Neural Network for Single-Molecule Nanopore Sensing," Nano Letters 18, 4040 -- 4045 (2018). 18A. Gelman, J. B. Carlin, H. S. Stern, and D. B. Rubing, Bayesian data analysis (Chapman & Hall, London, 1995). 19K. E. Hines, "A Primer on Bayesian Inference for Biophysical Systems," Biophysical Journal 108, 2103 -- 2113 (2015). 20D. S. Sivia and J. Skilling, Data analysis: A Bayesian tutorial (Oxford University Press, Oxford, 2006). 21D. J. MacKay, Information Theory, Inference and Learning Al- gorithms (Cambridge University Press, Cambridge, 2003). 22J. Canton, "LII. An essay towards solving a problem in the doctrine of chances. By the late Rev. Mr. Bayes, F. R. S. commu- nicated by Mr. Price, in a letter to John Canton, A. M. F. R. S," Philosophical Transactions of the Royal Society of London 53, 370 -- 418 (1763). 23N. Ermann, N. Hanikel, V. Wang, K. Chen, N. E. Weck- man, and U. F. Keyser, "Promoting single-file DNA translocations through nanopores using electro-osmotic flow," The Journal of Chemical Physics 149, 163311 (2018). 24W. J. Handley, M. P. Hobson, next-generation and A. N. Lasenby, sampling," nested "POLYCHORD: Monthly Notices of the Royal Astronomical Society 453, 4385 -- 4399 (2015). 25F. Feroz, M. P. Hobson, Nest: ence Monthly Notices of the Royal Astronomical Society 398, 1601 -- 1614 (2009). and cosmology Bayesian efficient and M. Bridges, "Multi- infer- physics," robust and particle tool for an
1212.1667
1
1212
2012-12-07T17:53:50
Microcanonical thermostatistics of coarse-grained proteins with amyloidogenic propensity
[ "physics.bio-ph", "physics.comp-ph", "q-bio.BM" ]
The formation of fibrillar aggregates seems to be a common characteristic of polypeptide chains, although the observation of these aggregates may depend on appropriate experimental conditions. Partially folded intermediates seem to have an important role in the generation of protein aggregates, and a mechanism for this fibril formation considers that these intermediates also correspond to metastable states with respect to the fibrillar ones. Here, using a coarse-grained (CG) off-lattice model, we carry out a comparative analysis of the thermodynamic aspects characterizing the folding transition with respect to the propensity for aggregation of four different systems: two isoforms of the amyloid $\beta$-protein, the Src SH3 domain, and the human prion proteins (hPrP). Microcanonical analysis of the data obtained from replica exchange method (REM) is conducted to evaluate the free-energy barrier and latent heat in these models. The simulations of the amyloid $\beta$ isoforms and Src SH3 domain indicated that the folding process described by this CG model is related to a negative specific heat, a phenomenon that can only be verified in the microcanonical ensemble in first-order phase transitions. The CG simulation of the hPrP heteropolymer yielded a continuous folding transition. The absence of a free-energy barrier and latent heat favors the presence of partially unfolded conformations, and in this context, this thermodynamic aspect could explain the reason why the hPrP heteropolymer is more aggregation-prone than the other heteropolymers considered in this study. We introduced the hydrophobic radius of gyration as an order parameter and found that it can be used to obtain reliable information about the hydrophobic packing and the transition temperatures in the folding process.
physics.bio-ph
physics
Microcanonical thermostatistics of coarse-grained proteins with amyloidogenic propensity Rafael B. Frigori,1, ∗ Leandro G. Rizzi,1, † and Nelson A. Alves1, ‡ 1Departamento de F´ısica, FFCLRP, Universidade de Sao Paulo, Avenida Bandeirantes, 3900. 14040-901, Ribeirao Preto, SP, Brazil. (Dated: October 12, 2018) Abstract 2 1 0 2 c e D 7 ] h p - o i b . s c i s y h p [ 1 v 7 6 6 1 . 2 1 2 1 : v i X r a The formation of fibrillar aggregates seems to be a common characteristic of polypeptide chains, although the observation of these aggregates may depend on appropriate experimental conditions. Partially folded intermediates seem to have an important role in the generation of protein aggregates, and a mechanism for this fibril formation considers that these intermediates also correspond to metastable states with respect to the fibrillar ones. Here, using a coarse-grained (CG) off-lattice model, we carry out a comparative analysis of the thermodynamic aspects characterizing the folding transition with respect to the propensity for aggregation of four different systems: two isoforms of the amyloid β-protein, the Src SH3 domain, and the human prion proteins (hPrP). Microcanonical analysis of the data obtained from replica exchange method (REM) is conducted to evaluate the free-energy barrier and latent heat in these models. The simulations of the amyloid β isoforms and Src SH3 domain indicated that the folding process described by this CG model is related to a negative specific heat, a phenomenon that can only be verified in the microcanonical ensemble in first-order phase transitions. The CG simulation of the hPrP heteropolymer yielded a continuous folding transition. The absence of a free-energy barrier and latent heat favors the presence of partially unfolded conformations, and in this context, this thermodynamic aspect could explain the reason why the hPrP heteropolymer is more aggregation-prone than the other heteropolymers considered in this study. We introduced the hydrophobic radius of gyration as an order parameter and found that it can be used to obtain reliable information about the hydrophobic packing and the transition temperatures in the folding process. PACS numbers: 87.15.Aa,87.14.Ee,87.15.Cc,05.70.Fh Keywords: amyloid β-peptides, Src SH3 domain, human prion protein, AB model, microcanonical analysis. I. INTRODUCTION Different diseases, including Alzheimer's [1, 2] (AD), Parkinson's [3], and Huntington's [4, 5] diseases, are known to be the result of neurodegenerative processes caused by formation of fibrillar aggregates of proteins, known as proteinopathies, which occurs primarily in the cellular structures [6 -- 10]. Because a number of proteins have been found to form fibrillar aggregates without sequence identity or structural homology under appropriate conditions in vitro, one can conclude that this is a common characteristic of polypeptide chains [7, 10, 11], that would reflect a property of their main backbone. Therefore, the use of detailed interatomic potentials in computational experiments employed for the thermodynamic characterization of proteins with known aggregation propensity seems to be inessential and thus, some general features can be extracted with the aid of simpler models [12]. In particular, we followed ref. [13] and utilized a coarse-grained off-lattice model [14] in our simulations. In the simulations, the 20 naturally oc- curring amino acid residues are replaced with monomers whose hydrophobic-polar characters are denoted by A and B, respectively. Thus, our investigation of the ther- modynamic features of peptide chains with known amy- ∗ Present address: Universidade Tecnol´ogica Federal do Paran´a, Toledo, PR, Brazil; [email protected][email protected][email protected] loidogenic propensity assumes that the nonbonded inter- actions are described by the hydrophobic-polar character of these monomers in all extension. The hydrophobicity of the side chains correlates with the aggregation rate [15, 16], so it is an important physicochemical ingredient promoting the nucleation of fibrillar aggregates. The mi- crocanonical analysis of the heteropolymer aggregation process has been performed [13, 17] and supports the main idea that simple models encompassing hydrophobic interactions among monomers or segments present the overall aggregation phenomenon through a temperature- driven first-order transition. A more comprehensive study on the properties of the aggregation phenomenon shows that the change in the aggregation rate depends on diverse factors, grouped into extrinsic or intrinsic factors [15, 18]. Extrinsic factors include physicochemical prop- erties related to the polypeptide environment. The in- trinsic factors associated with amyloid formation refer to the characteristics of the polypeptide chains. Factors like hydrophobicity, propensity to form the α-helical struc- ture, propensity to form the β-sheet structure, overall charge, and patterns of polar and nonpolar residues have been demonstrated to influence the aggregation rate. The propensity of globular proteins to adopt a different secondary structure content is another intrinsic factor related to the stability of the polypeptide chain. There is strong evidence that partially folded intermediates, such as for the Aβ peptide in the case of Alzheimer's, or with substantial unstructured conformations play an important role in the formation of amyloid fibril [19 -- 22]. Thus, conformational destabilization seems to be 1 a natural requirement for polypeptide chains to achieve new arrangements even under physiological conditions, which culminates in their assembly of amyloid fibrils [23]. On the other hand, unstructured or destabilized conformations do not seem to be a necessary condition to promote aggregation [24, 25], although results from mu- tations on the amyloid β-protein (Aβ) have shown that the less stable Aβ variants have the fastest nucleation kinetics in fibrillization [26, 27]. Here, thermodynamic features of the amyloid β- peptides (Aβ40 and Aβ42), the Src SH3 domain, and the human prion protein (hPrP) are compared in silico. Moreover, we try to associate those features with their aggregation propensity levels. The amyloid β-protein is the principal component of the amyloid plaques found in AD patients [28]. It is a small peptide that is predominantly constituted by 40 or 42 residues. The Src SH3 domain in turn, is a globular domain with 56 residues. Although this domain is non-pathogenic, it is expected to behave quite similarly to prion proteins during the aggregation transition. Actually, it has been experimentally shown that the SH3 domain aggregates to form amyloid fibrils [8, 29]. Molecular dynamics simulation of Src SH3 domains also led to aggregation conformations when the amino acid residue interactions were modeled by the G¯o potential [30]. The prion protein with PDB entry 1HJM [31] is considered in this study. A large number of observations have led to the conclusion that partial unfolding of the monomeric cellular prion protein (PrPC) seems to be necessary for the generation of the misfolded intermediates during fibrillization of the scrapie isoform PrPSc [24, 32, 33]. In addition, we emphasize that thermodynamic equi- librium properties are better analyzed in the microcanon- ical ensemble because it leads to the correct entropic characterization [34] and sets a simple way of determining free-energy barriers. In fact, many systems present equilibrium properties in the microcanonical ensemble but do not have their equivalents in the canonical en- semble [35 -- 37]. This is a consequence of the noncon- cavity of the entropy function, which produces energy- dependent equilibrium states in the microcanonical en- semble that cannot be associated with any temperature- dependent equilibrium states in the canonical ensemble. These missed states are the nonequilibrium ones in the canonical ensemble, and they generally correspond to metastable or unstable states in this ensemble [38]. The most relevant feature that emerges from the nonconcavity of the entropy is the occurrence of a negative micro- canonical specific heat. Hence, it is preferable to perform a microcanonical analysis so that the metastable states that may arise as a consequence of the dynamic structural conversion between the native and unfolded states can be properly accounted for. In this work, we use the replica- exchange method (REM) [39] and ST-WHAM-MUCA [40] analysis to obtain such microcanonical results and to base our discussions on. 2 II. MODEL AND SIMULATION ALGORITHM N−2(cid:88) (cid:32) N−2(cid:88) N(cid:88) To perform the simulations, the primary structure of a protein is mapped through the Roseman hydrophobicity scale [41] onto a string of elements A and B located at the Cα atoms. Table I includes the PDB codes used in this study and the corresponding AB sequences. The interaction for chains with N monomers is described by the following energy function [14], E = 1 4 (1 − cos θk) + 4 1 r12 ij − C (σi, σj) r6 ij k=1 i=1 j=i+2 (1) Here, θk is the angle between three successive monomers, and rij denotes the distance between the monomers i and j in the chain. The coupling constant C (σi, σj) in the Lennard-Jones type potential depends on the pairwise hydrophobic details, (cid:33) .  1, C (σi, σj) = (2) σi = σj = A σi = σj = B 0.5, −0.5, σi (cid:54)= σj . The spherical-cap algorithm [42] was used in the simula- tions in order to update conformations. This algorithm utilizes spherical coordinates for the generation of new positions for each monomer in the chain. Conformations at different temperatures were obtained by REM. Figure 1 presents a log plot of the temperatures used in the replica exchange simulations for the het- eropolymers Aβ, Src SH3, and hPrP. The temperature sets were determined by the following protocol: given an initial inverse temperature βn, we determined the next inverse temperature βn+1 by simulating only two replicas of the system with the Metropolis algorithm and the AB force field. First of all, we performed REM using Ns sweeps and Nswaps exchange conformation moves so as to equilibrate both replicas at the same reference temperature Tn = 1/βn, with the Boltzmann constant kB = 1. Next, the inverse temperature of one replica was increased by a small variation δβ, and Nswaps × Ns updates were performed again trying to exchange replicas after Ns sweeps. If the fraction of accepted replica exchanges facc was approximately equal to a probability pacc, this new inverse temperature βn+1 was accepted as reference; otherwise, the inverse temperature was increased by δβ once again. This procedure allowed for the recursive determination of all the temperatures. To compute facc, Nswaps = 50 and Ns = 2000 were employed. The acceptance probability pacc was set to 0.40 for the amyloid β-peptides, and 0.30 for the Src SH3 domain and the hPrP protein. It has been demonstrated that these pacc values provide a convenient number of round trips between extremal temperatures [43, 44]. For the all systems, evaluation of the temperature sets started at the inverse temperature β1 = 0.5, with δβ = 0.01. A simple statistics with five independent simulations was used for facc estimation, and thus for establishment of the set of temperatures {Ti} to perform REM simulations for each heteropolymer. Final data production was obtained with 12 replicas for Aβ40 and Aβ42, 16 replicas for Src SH3, and 20 replicas for hPrP. Simulations for data production were accomplished with Ns = 2000 sweeps and Nswaps = 10500 replica exchange moves, but for hPrP the number of exchange moves was doubled. Moreover, the above statistics was repeated five times, always starting from different initial conditions. Figure 2 shows conformations obtained from the AB model for the sequences in Table I, and sampled in the transition region. Data analysis was performed using the ST-WHAM procedure, which is an iteration-free approach that solves the WHAM equations in terms of intensive variables [45]. This procedure is a convenient way of obtaining the microcanonical inverse temperature estimates. On the basis of the assumption that the density of states can be obtained from Ω(E) ∝ H(E) W (E) , (3) with H(E) being the energy histogram and W (E) the simulation weight, ST-WHAM states that the inverse temperature can be estimated from M independent sim- ulations (n = 1,··· , M ), ∂ ln Ω(E) β(E) = ≈(cid:88) n ∂E f∗ n(βH n + βW n ), (4) (5) n = Hn/(cid:80) n Hn, βH n = ∂ lnHn/∂E, and βW where f∗ n = −∂ lnWn/∂E. The sum in n is over data produced by REM simulations, with the energy histograms Hn and Wn = e−E/Tn specified for each temperature Tn. Thermodynamic quantities such as microcanonical en- tropy S(E) and specific heat Cv(E) are evaluated from a multicanonical entropy-like solution [40], the so-called ST-WHAM-MUCA procedure. III. NUMERICAL SIMULATIONS AND RE- SULTS For comparative purposes, thermodynamic quantities such as the entropy, changes in the free energy ∆F , and latent heat (cid:96) are given as a function of the specific energy ε = E/N , where N stands for the total number of monomers in the system. These quantities are sum- marized in Table II. Because the force field treats the hydrophobic bonding interactions as the driving force in the folding process, it is important to analyze how the hydrophobic monomers behave as a function of the temperature. In fact, a recent study has shown that, besides hydrogen bonding, hydrophobicity is a fundamental interaction in the com- peting processes leading to folded or misfolded proteins 3 [46]. As we will show, the folding behavior can also be investigated by taking into account the spatial distribu- tion of hydrophobic monomers by means of a radius of gyration restricted only to this type of monomer. It has been demonstrated that this hydrophobic radius rh is a suitable scoring function for discrimination between the native structure and other conformations [47]. A. Microcanonical analysis of the Aβ models Figure 3(a) exhibits the estimates of β(ε) for the Aβ40 heteropolymer. The microcanonical entropy S(ε) is calculated using these ST-WHAM estimates of β(ε) according to the updating procedure ST-WHAM-MUCA. This entropy presents the so-called convex intruder (fig- ure not shown) and produces a negative specific heat in the energy range [−0.15,−0.07], which is a signature of a first-order phase transition. Figure 3(b) depicts the behavior of Cv(ε). A small positive peak in Cv(ε) is observed at ε = −0.579, presumably due to further com- paction of the heteropolymer chain. This latter transition is not related to a decrease in the microcanonical entropy with rising energy in this region. Therefore, β(ε) does not present a van der Waals-like curve in Fig. 3(b), which results in a continuous phase transition at βc = 2.695. To evaluate the free-energy profile as the heteropoly- mer assumes a folded conformation characterized by a stable phase within the metastable one, we considered the change of the microcanonical entropy between these phases in the vicinity of the transition temperature Tf = 1/βf . The energy region of interest is defined by the Maxwell construction in the caloric curve β(ε). This region is limited by the energies εa and εb (εa < εb) where the horizontal straight line of the Maxwell construction intercepts the caloric curve β(ε) and identifies the inverse of the canonical first-order transition temperature βf , which is related to the change in the entropy, βf = S(εb) − S(εa) εb − εa . 1 N (6) The energies εa and εb define the discontinuity or the latent heat, observed in van der Waals loops. To calculate the change in the free energy ∆F (ε) and βf from S(ε), we defined a shifted entropy between εa and εb, ∆S(ε) = N (A + βf ε) − S(ε), where A and βf are such that a canonical entropy, Scan(ε) = N (A + βf ε), is defined in that energy range. The constants A and βf are such that Scan(εa) = S(εa) and Scan(εb) = S(εb). These conditions yield εa = −0.191 and εb = −0.026, at Tf = 0.692(1) for Aβ40. The corresponding change in the free energy can be estimated as βf [F (ε) − Fa] = βf N ε − S(ε), (7) where βf Fa = βf N εa−S(εa) is the reference free energy. Our results for ∆F (ε) are shown in the inset of Fig. 3(a). This analysis yields the free-energy barrier ∆F = 0.038(2) at the folding temperature Tf . Latent heat is a consequence of the free-energy barrier that prevents the system from moving to a stable conformation in the new phase. Therefore, the smaller the latent heat, the higher the probability that a spontaneous fluctuation will give rise to this new phase. The estimate for the latent heat per monomer associated with this transition is (cid:96) = 0.165(3). A similar analysis follows for the heteropolymer ob- tained with the PDB entry 1Z0Q for the Aβ42 peptide. Results are presented in Fig. 4. However, for this het- eropolymer, the caloric curve displays a less pronounced van der Waals-like behavior around ε = −0.2 as a conse- quence of a minor nonconcavity of the microcanonical entropy. This model yields ∆F = 0.014(1) (inset of Fig. 4(a)) and latent heat (cid:96) = 0.125(2). Therefore, the free-energy barrier separating the folded and unfolded states of the heteropolymer describing Aβ42 is smaller compared with the Aβ40 heteropolymer. This corre- sponds to a less severe restriction to possible movements returning the heteropolymer to intermediate conforma- tions, depending on how stable the folded conformations are [48]. A small positive peak in the specific heat at ε = −0.686 signals a continuous transition (βc = 2.632) similar to the one observed for the Aβ40 heteropolymer. The results of the hydrophobic gyration radius rh for the Aβ40 and Aβ42 heteropolymers as a function of temperature are given in Fig. 5. The folding transition is expected to be accomplished by fast change in the spatial distribution of the monomers. In particular, this change is expected to be highly sensitive to the hydrophobic monomers, as illustrated in Fig. 5(a). Thus, we hypothesized that the temperature where the derivative < drh/dT > hits its maximum represents the occurrence of a thermodynamic transition. These temperatures, denoted by Tr, are easily identified in Fig. 5(b): Tr = 0.68(1) and Tr = 0.76(1) for Aβ40 and Aβ42, respec- tively. These figures also identify the latter compaction transitions at temperatures 0.37(1) and 0.38(1) for the Aβ40 and Aβ42, respectively. Estimates of Tr are in very good agreement with the values Tf obtained via Maxwell construction for both heteropolymers. This indicates that rh can be considered a reliable order parameter for depiction of a hydrophobic profile as a function of temperature. This conclusion is also supported by results attained for the Src SH3 and hPrP heteropolymers as shown in the following subsections. B. Microcanonical analysis of the Src SH3 model The SH3 domains have attracted much interest be- cause they represent typical examples of proteins that fold via a two-state mechanism. It is largely accept that the physical process underlying protein folding in these cases is based on the nucleation-condensation scenario [49 -- 52]. In fact, a sharp transition at the folding 4 temperature Tf between the unfolded and folded states has been observed [49, 50]. Results from simulations for the Src SH3 heteropoly- mer obtained with the AB model are shown in Fig. 6. As expected for such peptide with a clear two-state transition mechanism, a van der Waals-like loop can be detected in the caloric curve (Fig. 6(a)). Therefore, the microcanonical specific heat presents the typical behavior observed in a first-order phase transition (Fig. 6(b)). The Maxwell construction leads to εa = −0.182 and εb = 0.0079, with Tf = 0.658(1). For this model, we obtained a free-energy barrier ∆F = 0.068(2) (inset in Fig. 6(a)) and a latent heat per monomer (cid:96) = 0.190(8). Two small positive peaks are seen in Cv(ε) at ε = −0.21 and ε = −0.70. Both transitions are likely related to compactions. The behaviour of rh and its derivative are included in Fig. 5. The maximum of < drh/dT > occurs at Tr = 0.66(1). This feature also signals the existence of a second (minor) peak in the derivative of rh at Tr = 0.47(3), which is related to the transition observed in Fig. 6(a) at ε = −0.70. C. Microcanonical analysis of the hPrP model The next heteropolymer corresponds to the extracellu- lar globular domain hPrP with 104 residues. Compared to the previous models, the numerical results do not furnish any convex intruder in S(ε). The behavior of β(ε), displayed in Fig. 7(a), does not produce a van der Waals-like curve. The microcanonical analysis yields positive specific heats associated with continuous phase transitions (Fig. 7(b)). The maxima of Cv(ε) occur at ε1 = −0.36 and ε2 = −0.064, which correspond to the phase transition temperatures T1 = 0.596(2) and T2 = 0.697(1), respectively. The behaviour of rh and its derivative are also included in Fig. 5. Here, there is a clear transition at Tr = 0.60(1) and just a tiny change in the derivative of the hydrophobic radius at Tr = 0.32(1). Since our findings for this heteropolymer did not evidence a reduction in the microcanonical entropy as we moved toward the unfolded states, we may argue the following. Either the free-energy barrier separating the folded and unfolded states for this heteropolymer is not large enough to be revealed by our potential energy function because it does not contain interactions that reproduce the native contacts, or in fact the transition does not occur via a two-state mechanism. Interest- ingly, contradictory observations in what concerns the prion protein folding have been reported from diverse experiments. Experimental observations indicate that the native folding pathway involves the two-state process despite other evidences to the contrary (see, for example [33] and references therein). IV. DISCUSSION AND CONCLUSIONS The ability of peptide chains to form fibrillar ag- gregates seems to follow from their propensity to ag- gregate under conditions that permit partial unfolding. These fibrillar structures originate from misfolded protein chains via a complex process. A mechanism for the formation of the fibrils considers that the intermediates correspond to metastable states with respect to the fibrillar states. In this sense, the existence of these aggregation-prone states can be a consequence of the It has low degree of stability of the native states. been demonstrated with a simple lattice model [55] that low-energy conformations populate the aggregation- prone states. Such conformations are identified in the variety of lowest energy oligomers and protofilaments obtained when multiple chains are present in the three- dimensional lattice. Here, we performed REM simu- lations of four biologically inspired heteropolymers to relate the thermodynamic properties with their aggrega- tion propensities. Considering the amyloid β-peptides, there is experimental evidence that the Aβ42 is more aggregation-prone than Aβ40 [53]. Our findings for the free-energy barriers and latent heats of these het- eropolymer systems, as listed in Table II, indicate a two- state folding process. Smaller ∆F and (cid:96) for the Aβ42 heteropolymer, compared with the respective values for Aβ40, indicate that the formation of native states are facilitated for Aβ42. On the other hand, these numerical values signal a weaker first-order transition for Aβ42. This may give higher chances for the Aβ42 native-like conformations to adopt partially folded intermediates under similar stability conditions. As a matter of fact, monomeric Aβ peptides are considered to be intrinsically disordered and therefore, Aβ42 can cross the free-energy barrier to misfold conformations more easily [26]. For the Src SH3 heteropolymer, our results show a stronger first-order phase transition, confirming the two-state character expected for this system. Thus, on comparative grounds, it is reasonable to understand the experimental requirements to destabilise the native conformation [54] to produce some partially unfolded conformations as a prerequisite for its self-assembly. The conflicting evidence about the mechanism of the folding process involving PrP shows how difficult the experimental measurements are [33]. Interestingly, our in silico experiment suggests that if there is a free- energy barrier separating the folded and unfolded states of the hPrP heteropolymer, it is not large enough to be revealed by our simple potential energy function. We argue that the absence of a free-energy barrier favors the presence of partially unfolded conformations in the hPrP heteropolymer, which could explain why it is more aggregation-prone than the other heteropolymers. Of course, our conclusions are based on a simple force field which was not designed, for example, to reproduce the native state. However, we expect that the hydrophobic force incorporates the main aspects that produce the transition-state configurations. Because the propensity of peptides and proteins to form aggregates depends greatly on the sequence, Pawar et al. [18] defined a phenomenological equation to express this feature. We calculated the intrinsic score Zagg of aggregation for our polypeptide chains at pH 7 by means of the Zyggregator algorithm [56]. This algorithm yielded the values presented in Table II. Observation of the Z-scores reveals that the hPrP polypeptide is more aggregation-prone than the other polypeptide chains. This conclusion agrees with the thermodynamic phase transition results listed in Table II for these different heteropolymer systems, if we consider that weaker phase transitions facilitate the coexistence of mixed conforma- tions. 5, We also computed the hydrophobic radius of gyration and as can be seen from Fig. it is a convenient order parameter for identification of the folding tempera- tures. The agreement with the microcanonical estimates demonstrates that the spatial packing of hydrophobic monomers furnishes reliable information about the fold- ing process. More importantly, this quantity can easily be analyzed in any protein folding study and does not depend on any other information about the protein chain like the usual reaction coordinates needing some information about native contacts. ACKNOWLEDGMENTS: The authors acknowledge support by the Brazilian agencies FAPESP, CAPES, and CNPq. R.B.F. was also supported by UTFPR. This work used resources of the LCCA-Laboratory of Advanced Scientific Computation of the University of Sao Paulo. [1] D. J. Selkoe, Physiolog. Rev. 81, 741 (2001). [2] J. Hardy, and D. J. Selkoe, Science 297, 353 (2002). [3] M. R. Cookson, Ann. Rev. Biochem. 74, 29 (2005). [4] S. Ramaswamy, K. M. Shannon, and J. H. Kordower, 191 (2008). [6] S. Ohnishi, and K. Takano, CMLS, Cell. Molec. Life Sciences 61, 511 (2004). [7] F. Chiti, and C. M. Dobson, Ann. Rev. Biochem. 75, 333 Cell Transplantation 16, 301 (2007). (2006). [5] S. Imarisio, J. Carmichael, V. Korolchuk, C.-W. Chen, S. Saiki, C. Rose, G. Krishna, J. E. Davies, E. Itofi, B. R. Underwood, and D. C. Rubinsztein, Biochem. J. 412, [8] A. Espargar´o, V. Castillo, N. S. De Groot, and S. Ventura, J. Mol. Biol. 378, 1116 (2008). [9] R. M. Murphy, Annu. Rev. Biomed. Eng. 4, 155 (2002). 5 [10] D. Thirumalai, D. K. Klimov, and R. I. Dima, Curr. Opin. Struct. Biol. 13, 146 (2003). Cell 93, 337 (1998). Nature Rev. Genetics 6, 435 (2005). [11] M. Bucciantini, E. Giannoni, F. Chiti, F. Baroni, L. Formigli, J. Zurdo, N. Taddei, G. Ramponi, C. M. Dobson, and M. Stefan, Nature 416, 507 (2002). [37] M. Costeniuc, R. S. Ellis, H. Touchette, and B. Turking- ton, Phys. Rev. E 73, 026105 (2006). [38] H. Touchette, M. Costeniuc, R.S. Ellis, and B. Turking- ton, Physica A 365, 132 (2006). [39] K. Hukushima, and K. Nemoto. J. Phys. Soc. Jap. 65 1604 (1996). [12] B. Ma, and R. Nussinov, Curr. Opin. Chem. Biol. 10, [40] L. G. Rizzi, and N. A. Alves, J. Chem. Phys. 135, 141101 445 (2006). (2011). [13] C. Junghans, M. Bachmann, H. Arkin, and W. Janke, Phys. Rev. Lett. 97, 218103 (2006). [41] M.A. Roseman, J. Mol. Biol. 200, 513 (1988). [42] M. Bachmann, H. Arkin, and W. Janke, Phys. Rev. E [14] F. H. Stillinger, and T. Head-Gordon, Phys. Rev. E 52, 71, 031906 (2005). 2872 (1995). [15] K. F. DuBay, A. P. Pawar, F. Chiti, J. Zurdo, C. M. Dobson, and M. Vendruscolo, J. Mol. Biol. 341, 1317 (2004). [43] M. Lingenheil, R. Denschlag, G. Mathias, and P. Tavan, Chem. Phys. Lett. 478, 80 (2009). [44] C. E. Fiore, J. Chem. Phys. 135, 114107 (2011). [45] J. Kim, T. Keyes, and J. E. Straub, J. Chem. Phys. 135, [16] M. Belli, M. Ramazzotti, and F. Chiti, EMBO reports 061103 (2011). 12, 657 (2011). [17] T. Chen, X. Lin, Y. Liu, T. Lu, and H. Liang, Phys. Rev. E 78, 056101 (2008). [18] A. P. Pawar, K. F. DuBay, J. Zurdo, F. Chiti, M. Vendruscolo, and C. M. Dobson, J. Mol. Biol. 350, 379 (2005). [46] A. W. Fitzpatrick, T. P. J. Knowles, C. A. Waudby, M. Vendruscolo, and C. M. Dobson, PLoS Comput. Biol. 7, e1002169 (2011). [47] N. A. Alves, V. Aleksenko, and U. H. E. Hansmann, J. Phys.: Condens. Matt. 17, S1595 (2005). [48] D. Thirumalai, and G. Reddy, Nature Chem. 3, 910 [19] V. N. Uversky, and A. L. Fink, Biochim. Biophys. Acta (2011). 1698, 131 (2004). [20] D. Hamada, T. Tanaka, G. G. Tartaglia, A. Pawar, M. Vendruscolo, M. Kawamura, A. Tamura, N. Tanaka, and C. M. Dobson, J. Mol. Biol. 386, 878 (2009). [21] P. Neudecker, P. Robustelli, A. Cavalli, P. Walsh, P. Lundstrom, A. Zarrine-Afsar, S. Sharpe, M. Vendruscolo, and L. E. Kay, Science 336, 362 (2012). [22] J. R. Kumita, L. Helmfors, J. Williams, L. M. Luheshi, L. Menzer, M. Dumoulin, D. A. Lomas, D. C. Crowther, C. M. Dobson, and A.-C. Brorsson, FASEB J. 26, 192 (2012). [49] F. Ding, N. V. Dokholyan, S. V. Buldyrev, H. E. Stanley, and E. I. Shakhnovich, Biophys. J. 83, 3525 (2002). [50] J. M. Borreguero, N. V. Dokholyan, S. V. Buldyrev, E. I. Shakhnovich, and H. E. Stanley, J. Mol. Biol. 318, 863 (2002). [51] J. M. Borreguero, F. Ding, S. V. Buldyrev, H. E. Stanley, and N. V. Dokholyany, Biophys. J. 87, 521 (2004). [52] I. A. Hubner, K. A. Edmonds, and E. I. Shakhnovich, J. Mol. Biol. 349, 424 (2005). [53] J. D. Harper, and P. T. Lansbury, Annu. Rev. Biochem. 66, 385 (1997). [23] J. E. Straub, and D. Thirumalai, Annu. Rev. Phys. [54] Z. Liu, G. Reddy, and D. Thirumalai, J. Phys. Chem. B Chem. 62, 437 (2011). 116 6707 (2012). [56] The [55] M. S. Li, N. T. Co, G. Reddy, C.-K. Hu, J. E. Straub, and D. Thirumalai, Phys. Rev. Lett. 105, 218101 (2010). score http://www- intrinsic algorithm Zagg vendruscolo.ch.cam.ac.uk/zyggregator.php. propensity aggregation is available at [24] S. Liemann, and R. Glockshuber, Biochem. 38, 3258 (1999). [25] G. Soldi, F. Bemporad, S. Torrassa, A. Relini, M. Ramazzotti, N. Taddei, and F. Chiti, Biophys. J. 89, 4234 (2005). [26] C.-L. Ni, H.-P. Shi, H.-M. Yu, Y.-C. Chang, and Y.-R. Chen, FASEB J. 25, 1390 (2011). [27] G. Bitan, M. D. Kirkitadze, A. Lomakin, S. S. Vollers, G. B. Benedek, and D. B. Teplow, Proc. Natl. Acad. Sci. U.S.A. 100, 330 (2003). [28] A. Rauk, Chem. Soc. Rev. 38, 2698 (2009). [29] P. P. Laureto, N. Taddei, E. Frarel, C. Capanni, S. Costantini, J. Zurdo, F. Chiti, C. M. Dobson, and A. Fontana, J. Mol. Biol. 334, 129 (2003). [30] F. Ding, N. V. Dokholyan, S. V. Buldyrev, H. E. Stanley, and E. I. Shakhnovich, J. Mol. Biol. 324, 851 (2002). [31] L. Calzolai, and R. Zahn, J. Biol. Chem. 278, 35592 (2003). [32] M. L. DeMarco, and V. Daggett, Proc. Natl. Acad. Sci. U.S.A. 101, 2293 (2004). [33] H. Yu, X. Liu, K. Neupane, A. N. Gupta, A. M. Brigley, A. Solanki, I. Sosova, and M. T. Woodside, Proc. Natl. Acad. Sci. U.S.A. 109, 5283 (2012). [34] D. H. E. Gross, and J. F. Kenney, J. Chem. Phys. 122, 224111 (2005). [35] J. Barr´e, D. Mukamel, and S. Ruffo, Phys. Rev. Lett. 87, 030601 (2001). [36] F. Bouchet, and J. Barr´e, J. Stat. Phys. 118, 1073 (2005). 6 FIG. 1. Temperatures used to perform REM simulations. FIG. 2. Illustrative sampling of conformations for the AB sequences representing (a) Aβ40, (b) Aβ42, (c) Src SH3, and (d) hPrP proteins in the transition region. Dark (in color: red) color indicates hydrophobic monomers. (These images were made with VMD software support). 7 2468101214161820n0.20.40.60.8112TnAβ40Aβ42Src SH3hPrP FIG. 3. (a) Microcanonical inverse temperature estimates β(ε) = 1/T (ε), and (b) microcanonical specific heat Cv(ε) = −β2/(∂β/∂ε) for the Aβ40 heteropolymer. The inset of figure (a) shows the free-energy changes constructed from S(ε) at βf . FIG. 4. heteropolymer. The inset of figure (a) shows the free-energy changes constructed from S(ε) at βf . (a) Microcanonical inverse temperature estimates β(ε), and (b) microcanonical specific heat Cv(ε) for the Aβ42 8 11.522.533.54β-0.20ε00.050.1∆F-1-0.8-0.6-0.4-0.20ε-24-1201224Cv(a)(b)11.522.533.54β-0.20ε00.050.1∆F-1-0.8-0.6-0.4-0.20ε-24-1201224Cv(a)(b) FIG. 5. Behavior of the hydrophobic radius of gyration (a) and its derivative (b) as a function of temperature. FIG. 6. heteropolymer. The inset of figure (a) shows the free-energy changes constructed from S(ε) at βf . (a) Microcanonical inverse temperature estimates β(ε), and (b) microcanonical specific heat Cv(ε) for the Src SH3 9 5101520<rh>0.40.60.81T020406080100<drh/dT>Aβ40Aβ42Src SH3hPrP(a)(b)11.522.53β-0.20ε00.050.1∆F-1-0.8-0.6-0.4-0.20ε-24-1201224Cv(a)(b) FIG. 7. heteropolymer. (a) Microcanonical inverse temperature estimates β(ε), and (b) microcanonical specific heat Cv(ε) for the hPrP Protein PDB code Residues Sequence TABLE I. PDB codes and Roseman mapping. amyloid β amyloid β Src SH3 2LFM 1Z0Q 1NLO 40 42 56 BABAB BBBBB BABBB BAAAA ABBAB BBBBA AABAA ABBAA BABAB BBBBB BABBB BAAAA ABBAB BBBBA AABAA ABBAA AA AAAAA BBBBB BABAB ABABB BBBAB AABBA BBBAA AABBA AABBA BBAAB prion protein 1HJM 104 ABBBA ABBAA BBAAA BABBB BBBBB BBBBA BBBAB BABBB AABBB BBBBB AABBA ABAAA BBBAA AAAAB BBBAA BABAB AABBA ABBAA AABBB BBBBA BBAAA B BBBB TABLE II. Comparative results for the heteropolymer models. Model Tf ∆F (cid:96) Tr Zagg score [56] Aβ40 Aβ42 0.692(1) 0.769(1) Src SH3 0.658(1) 0.596(2) hPrP 0.038(2) 0.014(1) 0.068(2) 0.165(3) 0.125(2) 0.190(8) 0.68(1) 0.76(1) 0.66(1) 0.60(1) 0.90 0.94 0.96 1.14 10 11.522.53β-1-0.8-0.6-0.4-0.20ε0612Cv(a)(b)
1708.00190
1
1708
2017-08-01T07:30:56
Super-Gaussian, super-diffusive transport of multi-mode active matter
[ "physics.bio-ph", "cond-mat.soft" ]
Living cells exhibit multi-mode transport that switches between an active, self-propelled motion and a seemingly passive, random motion. Cellular decision-making over transport mode switching is a stochastic process that depends on the dynamics of the intracellular chemical network regulating the cell migration process. Here, we propose a theory and an exactly solvable model of multi-mode active matter. Our exact model study shows that the reversible transition between a passive mode and an active mode is the origin of the anomalous, super-Gaussian transport dynamics, which has been observed in various experiments for multi-mode active matter. We also present the generalization of our model to encompass complex multi-mode matter with arbitrary internal state chemical dynamics and internal state dependent transport dynamics.
physics.bio-ph
physics
Super-Gaussian, super-diffusive transport of multi-mode active matter Seungsoo Hahn1,2, Sanggeun Song1-3, Dae Hyun Kim1-3, Gil-Suk Yang1,3, Kang Taek Lee4* Jaeyoung Sung1-3* 1Creative Research Initiative Center for Chemical Dynamics in Living Cells, Chung-Ang University, Seoul 06974, Korea. 2Department of Chemistry, Chung-Ang University, Seoul 06974, Korea. 3National Institute of Innovative Functional Imaging, Chung-Ang University, Seoul 06974, Korea. 4Department of Chemistry, Gwangju Institute of Science and Technology, Gwangju 61005, Korea Corresponding Authors: J. Sung ([email protected]); K. T. Lee ([email protected]) Abstract Living cells exhibit multi-mode transport that switches between an active, self-propelled motion and a seemingly passive, random motion. Cellular decision-making over transport mode switching is a stochastic process that depends on the dynamics of the intracellular chemical network regulating the cell migration process. Here, we propose a theory and an exactly solvable model of multi-mode active matter. Our exact model study shows that the reversible transition between a passive mode and an active mode is the origin of the anomalous, super-Gaussian transport dynamics, which has been observed in various experiments for multi- mode active matter. We also present the generalization of our model to encompass complex multi-mode matter with arbitrary internal state chemical dynamics and internal state dependent transport dynamics. 1 Living cells in migration regulate their consumption of intracellular chemical energy according to the instructions encoded in their genes; they exhibit multiple transport modes during transport, consisting of two characteristic motions: a self-propelled, ballistic motion when the matter is in an active mode and an undirected, random motion when the matter is in a passive mode. Depending on the regulatory state of the cellular reaction networks underlying cell migration, the transport mode of living cells switches repeatedly between the active and the passive mode. This feature in living cell trajectories appears similar to that of Lévy walks [1,2]. Another interesting feature of living cells' motion is that they repeatedly reverse their direction. This run-and-reverse motion has been reported in various bacterial systems [3-7]. These features have also been observed in the transport of various types of cargos and vesicles in living cells [8-10]. Active matter, such as living cells and intracellular active particles, generally exhibits an anomalous, non-Gaussian transport dynamics, which cannot be described by Einstein's theory of Brownian motion [11,12] or more recent theories for anomalous transport in a disordered environment [13-17]. There are models of passively moving particles that have been used to explain the long time behavior of the mean square displacement (MSD) of multi-mode active matter observed in experiments [18,19]. Although these models assume that the stochastic dynamics of multi- mode active matter is qualitatively the same as that of passive matter, they are able to provide a satisfactory explanation of experimental results for the long time behavior of the MSD in many cases [20,21]. However, experimental data with a higher time resolution revealed that multi-mode active matter has qualitatively different stochastic dynamics from passive matter; the MSD of multi-mode active matter shows short time-diffusive motion, intermediate super- diffusive motion, and long-time diffusive motion with a greater diffusion coefficient [22,23], 2 which cannot be explained by the passive matter models [24,25]. An alternative model to account for the anomalous transport dynamics of active matter is the "active Brownian particle" model. In this model, a velocity-dependent friction in the Langevin equation is used to describe the self-propelled motion of active particles [7,26]. The active Brownian particle model does provide an enhanced explanation for the anomalous MSD of active particles; however, this model and the models mentioned above cannot explain the anomalous displacement distribution of active matter, whose spatial distribution is non-Gaussian with a positive excess kurtosis [6,27]. A number of other interesting models have been proposed for self-propelled particles [28-34]. However, to the best of our knowledge, none of them represents multi-mode active matter, which switches between an active, self-propelled transport mode and a seemingly passive, random mode depending on its internal state dynamics. In this Letter, we present an exactly solvable model for the stochastic transport of multi- mode active matter. In the high friction regime, where we can safely neglect the inertial term in the Langevin equation, the velocity ( )x t of multi-mode active matter with a friction constant, γ, can be written as the sum of two components: ( ) x t  = v s ( ( )) Γ t + 1 ( ) tγ ξ− , (1) where sv Γ and ( ) 1 ( )tγ ξ− represent the velocity component of a self-propelled, ballistic motion, which is dependent on the internal state Γ and the velocity component caused by the random fluctuating force. Assuming that the dynamics of the random fluctuating force occurs in a time scale far shorter than the internal state dynamics, we model ( )tξ as Gaussian white noise, whose time correlation, ( t ξ ) ( t ξ+ t 0 ) 0 , is proportional to the Dirac delta function, 3 ( )tδ . On the other hand, the relaxation of v s ( ( t Γ + t )) v s 0 ( ( t Γ 0 )) from the initial value, 2 sv , to the final value, 2 sv , occurs in the time scale of the internal state dynamics. We assume that the cell state dynamics is an arbitrary stochastic process that can be represented by a multidimensional Markov process. The Fokker-Planck equation corresponding to Eq. (1) is given by ∂ t ∂ P ( Γ , , x t ) = ∂ x ∂    D 0 ∂ x ∂ − v s ( Γ )    P ( Γ , , x t ) + L ( Γ ) P ( Γ , , x t ) , (2) where P ( Γ , , x t ) denotes the probability density function (PDF) of active matter with the position x and internal state Γ at time t [35,36]. In Eq. (2), 0D stands for the diffusion coefficient for passive motion originating from the random fluctuating force, which is defined by D 0 2 −= γ ∞ ∫ 0 dt ξ ξ ( ) (0) t . ( )L Γ denotes the mathematical operator describing the internal state dynamics of the system. Our model yields analytic results for the MSD and the non- Gaussian parameter [37]. Here, we compare two simple, exactly solvable models of active matter: one for single- mode active matter and the other for multi-mode active matter. These models are shown in Fig. 1. For the single-mode model, which only exhibits an active mode, as shown in Fig. 1(a), Eq. (2) yields ∂ t ∂ , P x t + , P x t − ( ( ) )         = ∂ x ∂    ∂ x ∂ D 0 I − v a 0    0 v − a           , P x t + , P x t − ( ( ) )     +    a k − k a k a k − a        , P x t + , P x t − ( ( ) )     . (3) For the multi-mode model, which exhibits both active and passive modes, as shown in Fig. 1(b), Eq. (2) yields 4 ∂ t ∂ , P x t + , P x t 0 , P x t − ( ( ( ) ) )           = ∂ x ∂      ∂ x ∂ D 0 I − v a 0 0      0 0 0 0 0 v − a                , P x t + , P x t 0 , P x t − ( ( ( ) ) )      + 0 k − k 0 0      k a 2 k − k a a 0 k 0 k − 0           , P x t + , P x t 0 , P x t − ( ( ( ) ) )      . (4) In Eqs. (3) and (4), ), iP x t ( designates the probability density of the matter at state iΓ ( , i ∈ + − ,0) and position x at time t. av± , ak , and 0k denote, respectively, the velocity sv ( ±Γ of the self-propelled motion of the matter at state ) ±Γ , the transition rate to either state, +Γ or −Γ , of the matter in active mode, and the transition rate to the state, 0Γ , of the matter in passive mode, at which 0( sv Γ = . Typical time traces are displayed for the two different ) 0 models in Fig. 1 and Supplemental Material [38]. Exact analytic solutions of Eqs. (3) and (4) can be obtained in the Fourier domain [37]. From the exact solutions, we obtain the distribution ( , ) f v t of the mean velocity, defined by ( ) x t ( t ≡ ( ) v t ) ; it is shown in Fig. 2 for each model. In the short-time limit, for both models, the mean velocity distribution is found to be a linear combination of Gaussians centered at the state-dependent self-propelled velocity, v Γ , that is, s ( ) i ( , ) f v t ≅ ∑ i ( eq p G v i − v s ( Γ i ), 2 D t 0 ) ( )c t τ<< , (5) with eq ip being the equilibrium probability of state iΓ , given by eqp± = 1/ 2 for the single- mode model and by eq p ± = k a 0( k + 2 ) k a and eq p 0 = k 0 0( k + 2 ) k a for the multi-mode model. In Eq. (5), G ν σ denotes the Gaussian distribution of ν with the mean and variance ( , ) 2 given by 0 and 2σ , respectively. Equation (5) can also be obtained from the distribution of the 5 instantaneous velocity given in Eq. (1), because the mean velocity, ( ) / x t t , is the same as the instantaneous velocity in the short-time limit [37]. Equation (5) serves as a good approximation of the mean velocity distribution at moderately short times, where a transition between states has yet to occur, or at times earlier than the characteristic relaxation time, ∞ ∫ 0 dt ( ) tφ sv [ ≡ τ c ] , where tφ denotes the normalized time correlation function, sv ( ) ( ) v t v s s (0) 2 v s tφ ( ) ≡ v s   , of the internal state dependent self-propelled velocity, ( ) sv Γ . The long-time distribution of the mean velocity becomes Gaussian for both single-mode and multi-mode models of active matter, while the short-time distribution can vary depending on the model. At short times ( t τ ), the mean velocity distribution, c ( , ) Sf v t , of the single- mode model has two Gaussian peaks centered at av+ and av− , which are the two velocities of self-propelled motion. In comparison, the mean velocity distribution, Mf ( , ) v t , of the multi- mode model at short times has an additional Gaussian peak centered at 0, resulting from the state, 0Γ , of the active matter in passive mode. The variance of each Gaussian peak, which originates from the random fluctuating force, is approximately given by 02D t . However, as shown in Fig. 2, both Mf ( , ) v t and ( , ) Sf v t converge to a Gaussian with a mean of zero and a variance proportional to 1/2 t − at long times [37]. That is to say, for both models, the distribution of ( )x t t approaches a Gaussian stable distribution at long times, in accordance with the Gaussian central limit theorem [37]. The relaxation dynamics of the mean velocity distribution is highly dependent on the 6 characteristic relaxation time, cτ , of internal state dependent self-propelled velocity, ( ) sv Γ . The mean velocity distribution approaches the long-time asymptotic Gaussian faster as the value of cτ decreases (see Figs. 2(c) and 2(d)). The analytic expression of cτ is dependent on the model in question. For the single-mode model, cτ is given by half the lifetime, 1 ak − , of the state, ±Γ , of the active matter in active mode, i.e., τ c = (2 ) ak 1 − . For the multi-mode model, cτ is the same as the lifetime, 1 0k − , of the state ±Γ [37]. In the small cτ limit, the mean velocity distribution is Gaussian at any finite time. Note also that the variance in the mean velocity, or the mean squared velocity, at any given time decreases with the relaxation speed, 1 cτ− , of the fluctuation in the self-propelled velocity, as shown in Fig. 2(e). This is a common feature of dynamically disordered systems; in spectroscopy, it has been termed motional narrowing. The mean velocity distribution, Mf ( , ) v t , of the multi-mode model is dependent on the lifetime, ( τ ≡ 0 1 2 ak ) , of passive mode, 0Γ , as well as on the lifetime, τ a ( ≡ 01 k = τ c ) , of the states in active mode, ±Γ . As shown in Fig. 2(f), when the population ratio, ( ≡ R eq p 0 ( eq p + + eq p − ) = τ τ a 0 = k 0 (2 ) a k ) , of the state in passive mode to the states in active mode decreases, Mf ( , ) v t approaches ( , ) Sf v t [37]. However, as the value of R increases, the peak centered at v = in 0 Mf ( , ) v t grows large, so that the MSD of the multi-mode model is smaller than the MSD of the single-mode model. For both the models, the MSD has three different kinetic phases: the short-time diffusion 7 phase, an intermediate super-diffusive phase, and the long-time diffusive phase with a greater diffusion coefficient, in agreement with the previous experimental results [22,23]. Exact analytic expressions of the MSD for both models can be written in the same formula, 2 x ( ) t = 2 ( ) D D t + 0 a + 2 ( τ − e a c D t ) τ 1c − , (6) where aD is the effective diffusion coefficient component contributed from the self-propelled motion, defined by D a ∞≡ ∫ 0 ( ) dt v t v s s (0) 2 v pτ = c a a . Here, ap designates the probability of the states in active mode, which is given by unity for the single-mode model and by ap = eq p + + eq p − (1 = + 1 ) R − for the multi-mode model. 0D and cτ have the same meaning as above. As shown in Fig. 3(a), the MSD is given by 2 x ( ) t ≅ 2 D t 0 at short times ( t τ ) and c dominantly contributed from the seemingly passive, random motion. On the other hand, at long times ( t τ ), the MSD is given by c 2 x ( ) t ≅ 2( D 0 + )a D t , with the diffusion coefficient increased by aD . In intermediate times ( 2 02 D v a t τ<  ), the MSD shows a super-diffusive c behavior ( MSD tα with 1 2α< ≤ ). In the early stage of the intermediate region, the MSD is approximately a quadratic function of time, i.e., 2 x ( ) t ≅ 02 D t D + τ τ a c c t ( )2 , shown by the green lines in Fig. 3(b), which originates from the ballistic, self-propelled motion of active matter. While both the single-mode and multi-mode models yield qualitatively the same analytic result for the MSD, the results they yield for the displacement distribution can be quite different from each other. The displacement distribution, ( , ) MP x t , of the multi-mode model can be 8 super-Gaussian, in accordance with the experimental data reported in Refs. [34,39], whereas ( , ) SP x t of the single-mode model is always sub-Gaussian. For the multi-mode model, the deviation of ( , ) MP x t of the multi-mode model from Gaussian measured by the non-Gaussian parameter, α  ( ) R t   ≡ 4 ( ) x t ( 3 2 ( ) x t )2 −  1   , is sensitive to the population ratio, R , of the state in passive mode to the states in active mode, which is shown in Fig. 3(c). The exact analytic expression of R tα is presented in the Supplemental Material [37]. The simpler ( ) asymptotic expression of R tα at both short times and long times is given by ( ) α R ( ) t   ≅     1 12 2 ( R ( R R ( ) 2 − ) 1 + R − − ) 3 1 R +     1   2 2 (0) D a D 0 2    (0) D a + D D a 0 ( t τ c 2 ) , t  τ c , t  τ c , 2    τ c t , (7) where (0) aD designates 2 c avτ , or the value of aD in the limit where the state of active matter is always in the active mode. According to Eq. (7), the displacement distribution, ( , ) MP x t , of multi-mode active matter is super-Gaussian when at all times [37]. However, when ( 1 ( 1 R < + 2 1.62 ) 5 ≅ 2R > , but sub-Gaussian when + ) 5 2 < 2R < , the displacement distribution, ( , ) MP x t , of multi-mode active matter switches from sub-Gaussian to super- Gaussian over time [37]. Note that R tα vanishes in the large R limit, where the state of ( ) multi-mode matter is always in passive mode. This means that, in our model, it is the self- propelled, ballistic motion that causes the displacement distribution to be non-Gaussian. In the opposite, small R limit, ( , ) MP x t has exactly the same shape as ( , ) SP x t [37]. Thus, the 9 non-Gaussian parameter, R tα , of the multi-mode model reduces to ( ) α 0 ( ) t  = the single-mode model, whose asymptotic behavior is given by lim ( ) t R → α R 0   of α 0 ( ) t    ≅      1 −  6     2 − 2  (0) D a  D  0 (0) D a + ( t τ c 2 ) , t  τ c . (8) 2    τ c t ) , t  τ c . (0 D D a 0 Equations (8) shows that the displacement distribution, ( , ) SP x t , of single-mode active matter is sub-Gaussian only [37]. Both ( , ) MP x t and ( , ) SP x t approach Gaussian at long times; however, their deviation from Gaussian, which is measured by the non-Gaussian parameter, slowly decreases with time, following 1t − at long times ( t τ ), according to Eqs. (7) and (8). As shown in Fig. 3(c), the c deviation of the displacement distribution from Gaussian can be sizable even at long times where the MSD, given in Eq. (6), is linearly proportional to time. This has been observed, for example, in liposome diffusion in a nematic solution of actin filaments [40]. The multi-mode active matter model discussed above can be extended to a more complex model in the higher spatial dimension, d. For the generalized model, the stochastic differential equation corresponding to Eq. (1) is given by r  ( ) t = v s ( ( )) Γ t + ξ 1 ( ) tγ− , (9) where each bold symbol denotes the d-dimensional vector corresponding to each scalar quantity in Eq. (1). The general expression of the MSD obtained from Eq. (9) is given by 10 r ( t 2 ) = 2 t ∫ d d 0 ( ) τ τ τ φ τ D 0 1 − p − ) ( t ξ   + D 1 ( ) − τ φ τ a c v s , (10)   where 0D , pτ , aD , and cτ are, respectively, defined by D d = 0 1 − 2 − γ ∞ ∫ 0 dt ξ ( ) t ⋅ ξ (0) , τ p φ∞≡ ∫ dt ξ 0 ( ) t , 1 −= D d a ∞ ∫ 0 dt v s ( ) t ⋅ v s (0) , and τ c φ∞≡ ∫ dt v 0 ( ) t . Here, ( )tφx s denotes the normalized time correlation function, x ( ) t ⋅ x (0) 2 x (0) , of vector ( )tx . The functional form of ( ) φ τv s varies depending on the internal state dynamics and its coupling to the self- propelled velocity. Given that the relaxation time of random fluctuation force ( )tξ is far shorter than the observation time t, Eq. (10) reduces to r ( t ) 2 ≅ 2 dD t 0 + 2 dD 1 − τ a c t ∫ 0 ( ( ) d τ τ φ τ − ) t v s . This result is the generalization of equation (6) for multi-dimensional systems with arbitrary ( ) φ τv s ; it reduces to equation (6) for the one- dimensional model with ( ) φ τ v s = exp( − t τ c ) . In addition, the general expression of the non- Gaussian parameter can also be obtained from equation (9) as follows: α R ( ) t = 2 2 ( ) t r v s 22 r ( ) t α v s ( ) t . (11) Here ( ) s tαv is defined by ( ) 4 t vr s defined ( ) t α ≡ v s d + 2 d as 4 ( ) t r v s 2 2 ( ) t r v s − 1 with ( ) 2 t vr s and t t ∫ ∫ 0 0 d dτ 2 τ 1 v s ( τ 2 ) ⋅ v s ( τ 1 ) and t t t ∫ ∫ ∫ ∫ 0 0 0 0 t d d d d τ τ τ τ τ 4 v ( 3 2 1 4 s ) ⋅ v s ( τ τ 3 2 v ) ( s ) ⋅ v ( τ 1 ) s , respectively. The non-Gaussian parameter given in Eq. (11) vanishes in both the short time and the long time limits. At times 11 far shorter than the relaxation time scale, cτ , of the self-propelled velocity, ( 2 ( ) t vr s 2 r ( ) t )2 , and hence the non-Gaussian parameter given in Eq. (11), vanish [37]. On the other hand, in the long time limit, ( 2 ( ) t vr s 2 r ( ) t )2 approaches D D D+ a a ( 0 ) but ( ) s tαv , or the non-Gaussian parameter of the self-propelled displacement, t ∫ 0 ( ) dτ τ v s , vanishes because the distribution of the self-propelled displacement becomes Gaussian according to the Gaussian central limit theorem. However, the non-Gaussian parameter has a non-zero value between the two limits. In the simple one-dimensional multi-mode active matter model with the Poisson state switching dynamics, we can show that equation (11) reduces to equation (7) [37]. Equations (10) and (11) enable us to calculate the MSD and non-Gaussian parameter for general multi-mode active matter with possibly non-Poisson state switching dynamics. In summary, we present an analytic theory and an exactly solvable model of multi-mode active matter, which switches between an active, self-propelled transport mode and a seemingly passive, random mode depending on its internal state chemical dynamics. Our exact model study clearly shows that the reversible transition between seemingly passive, random motion and the self-propelled, ballistic motion is an important source of the super-Gaussian displacement distribution commonly observed for multi-mode active matter. This model is sufficiently flexible so that it can be easily generalized to encompass multi-state, multi-mode active matter with arbitrary internal state chemical dynamics and internal state coupled transport dynamics. The application of the present approach to the quantitative explanation of experimental results for examples of multi-mode active matter is to be published elsewhere. 12 FIGURES +Γ and −Γ . The single-mode active matter in FIG. 1. Model systems and typical trajectories. (a) The single-mode model consists of two ±Γ state performs self- internal states, av± under a random fluctuating force exerted from propelled, directed motion with velocity medium. The stochastic transition between internal states is characterized by the rate constant, ak . (b) The multi-mode model consisting of three internal states: passive transport state, 0Γ , −Γ . The multi-mode matter performs undirected, in addition to active transport states, av± random motion in state 0Γ to the in state 0Γ state, respectively. For each active model, a typical time trace of the position is shown. Colors in the active matter diagram and trajectory represent the cell's internal states. ±Γ . ±Γ state and from the active 0Γ , but performs directed, self-propelled motion with velocity ak and 0k represent the stochastic transition rates from the passive +Γ and ±Γ to the passive 13 uT , and the relaxation time for the single-mode model and (b) FIG. 2. PDFs for mean velocity distribution. The time dependent mean velocity distribution, ( , ) Sf v t (a) for the multi-mode model with 0.5 R = . In both (a) and (b), the mean velocity distribution is displayed starting from arbitrary cτ of the velocity-velocity auto-correlation function is unit time, set to be 10 uT . The mean velocity distribution at (c) for the single-mode model and (d) cτ . (lines) analytic results (circles) for the multi-mode model, with three different values of stochastic simulation results. In (d), the value of R is set to be 0.5, in which case the three states are equally probable at equilibrium. (e) Dependence of the root-mean-square velocities, or the standard deviation of the mean velocity distributions on the relaxation speed measured by for the multi-mode model with three ; (black line) different values of R : (blue dotted line) ( ,1) v R = ( ,1) v approaches 5R = is plotted as a red dotted line. The values of the other parameters are set to be for 1D = and 0 cτ is 10 uT . In the small R limit, . The Gaussian distribution with the same mean and variance as cτ− , and (f) the mean velocity distribution at 0R = ; (blue solid line) av = for all cases. 5 0.5 ; and (red line) 5R = . The value of t T= u t T= u ( ,1) Sf v 1 Mf ( , ) v t R = 0.05 Mf Mf 14 cτ = 10 2 cτ = 0.1 ≅ 02 D t D + 0D and , whose transition time scale is determined by (red), (blue). The effective diffusion coefficient increases from cτ is set to 1. The values of the other parameters, 10 aD = 2( ) 2 x t FIG. 3. Mean square displacement and non-Gaussian parameter. (a) Time-dependent mean square displacement (MSD). The single-mode and multi-mode models share the same MSD, aD given in Eq. (6). The value of 1D = and are set to be . (lines) analytic results (circles) stochastic simulation 0 1 cτ = results. (b) Dependence of t on time for the three cases with 0D to (black), and D D+ cτ . The green line represents the 0 a ( ) t x ) corresponding to each case. (c) The non- ballistic motion ( ( )tα , for the single-mode model (blue line) and for the multi-mode Gaussian parameter, M tα for the two models with various values of R (black lines). The two red lines represent ( critical values of R , 1.62 and 2. (d) , 1 2 1.62 R < + ( )M tα switches from a short-time but positive at all times when negative regime to a long-time positive regime. (e) As a representative case for the time- dependent switching, is plotted as a black line. The two red lines M tα for the two critical values of R , shown in (c). The non-Gaussian parameter represent vanishes both in the short time and the long time limits, meaning that the initial distribution is a delta function, Gaussian with zero variance, and the final distribution obeys the Gaussian central limit theorem. At the four time points marked by the solid circles, the mean velocity ( )M tα is always negative when 2R > . When 1.62 M tα with ( ) ( ) 5 τ τ a c c t ( )2 2R< < , ) ≅ R = 1.75 ( ) 15 distributions (blue line) and their corresponding Gaussian distributions (black line) are plotted according to the square mean velocity in the insets. The red circles mark the deficiency in population of the active matter in the high mean velocity region, compared to their corresponding Gaussian distribution. (f) Deviation of the mean velocity distribution from Gaussian at the two time points marked by the filled black circles. The red circles here and in the insets in (e) both represent the same spatial regime. 16 G. Ariel, A. Rabani, S. Benisty, J. D. Partridge, R. M. Harshey, and A. Be'er, Nat. Commun. 6, T. H. Harris et al., Nature 486, 545 (2012). J. E. Johansen, J. Pinhassi, N. Blackburn, U. L. Zweifel, and A. Hagstrom, Aquat. Microb. Ecol. O. Sliusarenko, J. Neu, D. R. Zusman, and G. Oster, Proc. Natl. Acad. Sci. U. S. A. 103, 1534 R. Grossmann, F. Peruani, and M. Bar, New J. Phys. 18, 043009 (2016). Y. Wu, A. D. Kaiser, Y. Jiang, and M. S. Alber, Proc. Natl. Acad. Sci. U. S. A. 106, 1222 (2009). P. Romanczuk, M. Bar, W. Ebeling, B. Lindner, and L. Schimansky-Geier, Eur. Phys. J. Spec. S. H. Nam, Y. M. Bae, Y. I. Park, J. H. Kim, H. M. Kim, J. S. Choi, K. T. Lee, T. Hyeon, and Y. D. K. Chen, B. Wang, and S. Granick, Nat. Mater. 14, 589 (2015). D. Arcizet, B. Meier, E. Sackmann, J. O. Radler, and D. Heinrich, Phys. Rev. Lett. 101, 248103 L. Schimanskygeier, M. Mieth, H. Rose, and H. Malchow, Phys. Lett. A 207, 140 (1995). References [1] 8396 (2015). [2] [3] 28, 229 (2002). [4] (2006). [5] [6] [7] Top. 202, 1 (2012). [8] Suh, Angew. Chem. Int. Ed. Engl. 50, 6093 (2011). [9] [10] (2008). [11] [12] 1954). [13] [14] [15] [16] 1994). [17] [18] [19] [20] [21] [22] [23] Lett. 99, 048102 (2007). [24] [25] C. Cox, and H. Flyvbjerg, Eur. Phys. J. Spec. Top. 157, 1 (2008). [26] 17 A. Einstein, Ann. Phys. 322, 549 (1905). N. Wax, Selected papers on noise and stochastic processes (Dover Publications, New York, E. W. Montroll and G. H. Weiss, J. Math. Phys. 6, 167 (1965). V. M. Kenkre, E. W. Montroll, and M. F. Shlesinger, J. Stat. Phys. 9, 44 (1973). J. Klafter and R. Silbey, Phys. Rev. Lett. 44, 55 (1980). G. H. Weiss, Aspects and applications of the random walk (North-Holland, Amsterdam, R. Metzler, E. Barkai, and J. Klafter, Phys. Rev. Lett. 82, 3563 (1999). C. L. Stokes, D. A. Lauffenburger, and S. K. Williams, J. Cell Sci. 99, 419 (1991). M. H. Gail and C. W. Boone, Biophys. J. 10, 980 (1970). L. Li, S. F. Norrelykke, and E. C. Cox, PLoS One 3, e2093 (2008). X. Liu, E. S. Welf, and J. M. Haugh, J. R. Soc. Interface 12, 20141412 (2015). A. J. Loosley, X. M. O'Brien, J. S. Reichner, and J. X. Tang, PLoS One 10, e0127425 (2015). J. R. Howse, R. A. Jones, A. J. Ryan, T. Gough, R. Vafabakhsh, and R. Golestanian, Phys. Rev. D. Campos, V. Mendez, and I. Llopis, J. Theor. Biol. 267, 526 (2010). D. Selmeczi, L. Li, L. I. I. Pedersen, S. F. Nrrelykke, P. H. Hagedorn, S. Mosler, N. B. Larsen, E. M. Theves, J. Taktikos, V. Zaburdaev, H. Stark, and C. Beta, Biophys. J. 105, 1915 (2013). D. Campos and V. Mendez, J. Chem. Phys. 130, 134711 (2009). N. Masaki, H. Miyoshi, and Y. Tsuchiya, Protoplasma 230, 69 (2007). H. Miyoshi, N. Masaki, and Y. Tsuchiya, Protoplasma 222, 175 (2003). F. Peruani and L. G. Morelli, Phys. Rev. Lett. 99, 010602 (2007). M. Schienbein and H. Gruler, Bull. Math. Biol. 55, 585 (1993). H. Takagi, M. J. Sato, T. Yanagida, and M. Ueda, PLoS One 3, e2648 (2008). D. Selmeczi, S. Mosler, P. H. Hagedorn, N. B. Larsen, and H. Flyvbjerg, Biophys. J. 89, 912 [27] [28] [29] [30] [31] [32] [33] [34] (2005). [35] [36] Verlag, New York, 1996), 2nd edn., Springer series in synergetics,, 18. [37] See Supplemental Material at for the derivation of the second and fourth moments of displacements, for PDF of displacements at limiting time scales, for mean velocity distriubtion and stationary distribution, for more details on the extended model, for the relaxation times, for non- Gaussian parameter in all range, and for the stochastic simulation method. [38] See Supplemental Material at for 200 trajectories of each model. H. U. Bodeker, C. Beta, T. D. Frank, and E. Bodenschatz, Europhys. Lett. 90, 28005 (2010). [39] B. Wang, J. Kuo, S. C. Bae, and S. Granick, Nat. Mater. 11, 481 (2012). [40] S. I. Denisov, W. Horsthemke, and P. Hanggi, Eur. Phys. J. B 68, 567 (2009). H. Risken, The Fokker-Planck equation : methods of solution and applications (Springer- 18 Supplemental Material for "Super-Gaussian, super-diffusive transport of multi-mode active matter" Seungsoo Hahn,2 Sanggeun Song,1,2,3 Dae Hyun Kim,1,2,3 Gil-Suk Yang, 1,2,3 Kang Taek Lee, 4 Jaeyoung Sung1,2,3,† 1Creative Research Initiative Center for Chemical Dynamics in Living Cells, Chung-Ang University, Seoul 06974, Korea. 2Department of Chemistry, Chung-Ang University, Seoul 06974, Korea. 3National Institute of Innovative Functional Imaging, Chung-Ang University, Seoul 06974, Korea. 4Department of Chemistry, Gwangju Institute of Science and Technology, Gwangju 61005, Korea Contents A. Derivation of the second and fourth moments of displacement for the multi-state model B. Derivation of the second and fourth moments of displacement for the single-state model C. Probability density function of displacement at two limiting time scales D. Mean velocity distribution and stationary distribution E. Dynamics of the multi-state model 2 4 5 7 8 8 9 10 11 12 13 14 15 ( , ) MP x t to ( , ) SP x t at the small R limit F. Convergence of G. Relaxation time of the two solvable models H. General model I. Time correlation functions for two solvable models J. Derivation of short time mean velocity distribution from Eq. (1) K. Stochastic simulation method L. References Fig. S1. Non-Gaussian parameter 1 A. Derivation of the second and fourth moments of displacement for the multi-mode model The multi-mode active matter model has three internal states, −Γ . Each internal state regulates the direction and speed of a given active matter as explained in Fig. 1(b). Based on the initial conditions that the internal states are initially in equilibrium and the initial position of active matter is zero, three simultaneous equations are obtained from Eq. (4) by applying the Fourier transform and the Laplace transform to − . The solution of the simultaneous equations provides three probability density functions (PDF) for the individual internal states in the Fourier-Laplace domain, written as 0Γ , and ), ( iP x t with +Γ , i ∈ + , 0,  , P w s +  , P w s 0  , P w s − ( ( (      with ) ) )      ( χ = ( s + 2 D 0 w ) ( ( χ 1 ) , w s + 2 2 v w a ) + 2 2 k v a a w , w s ) ≡ ( s + D 0 2 w k + 0 )( s D w k + + 2 0 + 0 2        2 a k eq p + ( ( χ , w s ) eq p 0 ) + i − ( i ( v w s a ( , w s χ ( v w s a + ) + + 2 w k D + 0 ) 2 v w a 2 w k + 2 ( χ , w s ) ( ) eq p − ) . (S1) D 0 2 + k 0 a + 2 k a 0 ) )        In Eq. (S1), w and s respectively denote the Fourier transform of position x and the Laplace transform of time t. The tildes indicate that the functions are represented in the Fourier-Laplace 0Γ , and domain. −Γ , respectively. A summation of the PDFs given in Eq. (S1) provides the PDF of multi-mode active matter, which is given by eqp− denote the equilibrium probabilities of the states eqp , and 0 eqp+ , +Γ ,  , P w s M ( ) ≡  , P w s − ( ) = ( s + D 0 w ( )  , P w s + 0 ( ) , w s + χ ( ) ) ( , w s χ 2 (  , P w s + + 2 2 eq p v w 0 a ) 2 2 v w + a + ) . (S2) 2 k 2 v a a 2 w + The denominator ( ) 3 k z z + a ) ( iC w − function as 2 + k k ( 2 0 in Eq. ( k k 0 with ( 2 + 0 a i ∈ ) is + cubic 2 a 2 k v w a a (S2) ) 2 2 v w z + a , the PDF can be rewritten as 1, 2,3) 2 function of as . If we assume the roots of the cubic ≡ + s z 2 wD 0 1 D 0  , P w s M ( ) = s = s + + 1 D 0 − 2 2 k v w a a 2 2 w − 2 2 k v w a a 2 2 w       s s + + 1 D 0 1 D 0 + 2 w k + k 0 2 w 3 ∏ 1 = 1 2 + 0 1 2 + k a k a    i    3 ∑ i 1 = 1 2 + C i ( s + D w 0 1 2 + w s + D 0 ( C w i ) ) w ∏ 3 j ≠ i 1 − ( C w C w i ( ) j . (S3) ) Inverting s in ),P w s (  generates the Fourier-domain PDF, written as , P w t M ( ) = 2 2 v w a e − D 0 w 2 t     3 ∑ i 1 =     2 k a ( C w i ) − 2 k a 2 + k a k 0 e     − ) C w t i ( 2 3 ∏ j ≠ i 1 − ( C w C w i ( ) j     ) . (S4) Eqs. (S2) and (S4) can both be used to derive the analytic solution for the mean square displacement (MSD) of the multi-mode model. One way is to use the second partial derivative , while the other way, which is an easier way to obtain the time-domain MSD, is of to apply the inverse Laplace transform to the second partial derivative of , written as , MP w t ),  MP w s ( ( ) 2 x ( ) t = lim 0 w → −     2 ∂ ) ( , P w t M w ∂ 2     1 − = L → t s     lim 0 w → −    2 ∂ ) (  , P w s M w ∂ 2        , (S5) where 1 L − t → s denote the inverse Laplace transform of time. The time-dependent MSD of this model is given in Eq. (6). The analytic solution for the fourth moment of displacement is obtained using the following equation: 4 x ( ) s = im l 0 w → 4 ∂    ) (  , P w s M w ∂ 4    = 24 3 s        2 D 0 + ( s + k     3 D 0 − s 2 2 v k a a )( k 0 1 k + 0    0 × + 2 k a ) 2 k D v − 0 a 0        + s + 2 v k 0 a k + 0 2 k a        . (S6) Application of the inverse Laplace transform to Eq. (S6) provides the fourth moment of displacement in the time domain, written as 4 x ( ) t = + 24 ( R 2 2 D τ a c ) 2 1 + e 3 ( R − 12         + + 5 ( ( ( ) 2 D D t + 0 2 a ) R t / τ c ( 1 1/ − + ( + − 6 − R − R 1 − + 2 R − 2 3 3 R + ) D D 0 a ( 2 R − − 2 − 3 R R R + + 2 R − ) ( 2 1 R ( ) 2 1 R )( 3 R + ) 3 1 e − t / τ c t − ) e ) 1 + / τ c t D 0 τ c D a ) t τ c (S7)         The second and fourth moments expressed in the time domain are combined to produce the non-Gaussian parameter, ( )tα , such as 3 α R ( ) t ≡ 3 4 x x ( ) t ( ) t 2 11 − = 3 2 κ ( ) t − 1 = ( R 1 + 4 ) 1     (0) 2 D τ a c ( ) 2 t x τ c − 2 t − t − 4e τ − c + t − 5 4e − τ + c τ c t − τ c − t 2 t τ c τ c 4 − t τ c ) R t τ c ) 3 R 2 − t τ c   −   + 2    +    +      e ( ( ( 2e − 2 − e − − 6e  2 e +   − t t τ c τ c 1 R t τ c − t τ c t τ c 8e 1 2e + − − 6 4e − + 1 R 1 − + t − τ c t 10 8e − ) t R τ c 2 t + τ c 2 t τ c    − t τ c e 5 R              with 2 ( ) t x ) (0 2 D τ a c = D t 0 (0) D τ a c + ( R  1  )  1 +  t − τ c E + t τ c − 1     , (S8) ) c ( + R ) 1 (0) a ( )tκ denotes kurtosis and (0 2 aD vτ D = ≡ where . On the log-scale time axis as c a a ( ) cτ, shifts the R tα curve as well as the MSD curve. shown in Fig. 3 (b), the relaxation time, ( ) t τ is analytically evaluated and plotted under the R tα in all ranges of R and In Fig. S1, D D = condition of , where the red lines are the two lines shown in Figs. 3(c) and 3(e) 0 and the black line marks a border line switching from sub-Gaussian to super-Gaussian at a D D , the border line is invariant on the change of given R. Although 0 ( ) R tα with R less than ( is sub-Gaussian at the 1 ( ) R tα with R larger than 2 is always super-Gaussian at all times. When can switch from sub-Gaussian to super-Gaussian over time, as < , 2R ( )tα depends on D D ratio in Eq. (S8). Thus, 0 0.1 ( ) 0 a ( ) 0 a ) 5 2 + + 5 2 ) all times, and ( ( , ) MP x t 1 shown in Figs. 3(e) and S1. < B. Derivation of the second and fourth moments of displacement for the single-mode model The single-mode model has two internal states, −Γ . Each internal state regulates the direction and speed of active matter, as explained in Fig. 1(a). Based on the initial conditions that the internal states are initially in equilibrium and the initial position of active matter is zero, two simultaneous equations are obtained from Eq. (3) by applying the Fourier transform and the Laplace transform to i ∈ + − . The analytic solution of the simultaneous equations provides two PDFs in the Fourier-Laplace domain, written as +Γ and ( ), iP x t with , 4      , P w s +  , P w s − ( ( ) )     = 1 2 2 2 v w a + ( 1 )( s D w s D w + + 2 0 0    ) 2 + 2 k a 0 s D w + s D w + 0 2 2 + + 2 2 k k a a − + i v w a i v w a    . (S9) The PDF of active matter for the single-mode model in Fourier-Laplace domain is written as  , P w s S ( ) ≡ (  , P w − s ) + s s D w s D w + (  , P w s + 2 D w + 0 )( 2 ) 2 + + k 0 0 a = 2 2 v w a + ( . (S10) 2 + 2 k a ) Application of the inverse Laplace transform to Eq. (S10) generates the PDF of active matter represented in the Fourier domain as , P w t S ( ) − = e ( t k D 0 + a 2 w )    cosh ( t ) Λ + k a Λ in s h ( t Λ )    with Λ ≡ k 2 a − 2 2 v w a . (S11) From this function, the time-dependent second and fourth moments of displacement are simple to obtain. The time-dependent MSD of this model is given in Eq. (6). The fourth moment of displacement is also evaluated from the PDF as 4 x ( ) t = 2 12 τ c     ( D D 0 a + 2 ) 2 t 2 τ c + 2 D D D 0 a − a (     )( 1 e − − t τ c ) t τ c + 3 D a    1 e − − t τ c − t τ c           , (S12) D a 2 vτ= c a where to produce the non-Gaussian parameter, ( )tα , such as . The second and fourth moments expressed in the time domain are combined α 0 ( ) t =     2 D τ a c ( ) 2 t x 2     ( 5 e − 2 − t τ c − 4e − t τ c − 4e − t τ c t τ c − 2 t τ c ) =     2 D τ a c ( ) 2 t x 2     β ( ) t with 2 ( ) x t 2 D τ a c = D t 0 D τ c a t − E τ c +    + t τ c − 1    and ( ) t β ≡ 5 e − 2 − t τ c − 4e − t τ c − t − 4e c τ τ c t − 2 t τ c . (S13) ( ) In Eq. (S13), equation is equal to ( ) 0 tβ ≤ in all time ranges, and the ( ) . t ( ) t = α 0 lim R α R → 0 tα is less than or equal to zero because 0 R tα of the multi-mode model at the small R limit as ( ) C. Probability density function of displacement at two limiting time scales The diffusion dynamics of the models is highly dependent on the relaxation time, cτ, of the t τ ), a given active matter maintains its direction and velocity, ( ) sv Γ . At short times ( c 5 magnitude of velocity, and each unrelaxed velocity produces three individual peaks in the PDF of displacement. The PDF of displacement at short times is derived from Eq. (S1), which is written as       P short  P short  P short , + ,0 , − ( ( ( , w s , w s , w s ) ) )      = eq p + e q p −        s D 0 ( eq p 0 ( + ( Ds + 0 s + 2 w + i 2 D 0 w 2 w − i v w a ) 1 − v a w 1 − ) 1 − )        . (S14) ), MP x t ( at short times is written as P M sho , rt ( , x t ) = 1 4 D t π 0     2 ) ( − x v t + a 4 D t 0 eq p e − − 2 x D t 0 4 + eq p e 0 + eq p e + ( − 2 ) t x v − a 4 D t 0     , (S15) ) , x t where the distribution is Gaussian with a variance of are approximated as a single Gaussian function with a small variance at very short times t ( ) and gradually separate as time increases. 02D t . The three peaks in P M short ( 2 02 D v a  , At long times ( t τ ), the peaks for individual c ( ) sv Γ s are again intermingled into a single Gaussian and follow the distribution, written as       P long  P long  P long , + , 0 , − ( ( ( , w s , w s , w s ) ) )      = s + 1 D eff 2 w       eq p 0 ( i1 eq p − + ( 2 2 1 v w k + 0 a ( i 1 eq v p a − ) v w k 0 a ( 2 k + 0 ) w k + 0 ) ) k a       , (S16) where effD is equal to D D+ 0 a . The PDF ), MP x t ( at long times is written as P M long , ( , x t ) = 2 x D t eff − 4 e 1 4 D tπ e f f 1 +     eq p D R 0 2 τ c a D t eff     1 − 2 x D t eff 2         . (S17) effD t , In Eq. (S17), deviation from the Gaussian distribution is proportional to t τ  at long times. eq p D D is always less than 1, and t is larger than where 0 Therefore, is much smaller than 1 if R is finite at sufficiently long times. Thus, the PDF approaches the Gaussian distribution at sufficiently long times, which is in accordance with the Gaussian central limit theorem. eq p D Rτ 0 a c 1 a eq p D Rτ 0 a c cτ as effD t 2 2 eff c In summary, the PDF of displacement approaches the delta function at very short times, because self-propelled velocity is much weaker than the velocity caused by the random fluctuating force. As a result of the contribution of random fluctuation being dissipated, the delta function is split into individual peaks related to the velocity of each internal state, whereas 6 the self-propelled velocity shows no variation. Eq. (S15) explains the two different functional forms of the PDF. At long times, the PDF approaches the dispersed Gaussian distribution because the variance for each distribution is proportional to time. D. Mean velocity distribution and stationary distribution The mean velocity, ), v t . The mean velocity distribution, , is directly obtained from the PDF of displacement with a proper normalization , is defined by ( )v t ( ) v t ( ) x t ≡ t ( Mf constant. At short times ( t τ ), the mean velocity distribution, c Mf ( ), v t , is written as f M short , ( , v t ) = 1 4 D t π 0     − 2 ( v 4 ) v + a D t 0 eq p e − − 2 v D t 0 4 + eq p e 0 + eq p e + − 2 ( v 4 ) v − a D t 0     . (S18) 02D t . Because the variance is inversely where the distribution is Gaussian with a variance of proportional to t, the broadness of the individual peaks in Eq. (S18) shows the opposite pattern compared to the individual peaks in , which appears in Eq. (S15). At long times t τ ), ( is written as ), MP x t ), v t ( ( Mf c f M long , ( , v t ) = 1 4 D t π ef f − 2 v D t eff 4 e 1 +     eq p D R 0 2 τ c a D t eff     1 − 2 v D eff 2         t . (S19) f M long , ( , v t ) approaches the delta function as time increases. times. If we define a new variable, The variance of the individual peaks in ), MP x t ) t at two the time-limiting cases. At short times ( written as ( ≡ q x c ( is proportional to t at both short and long , then the stationary distribution can be obtained t τ ), the stationary distribution, ), q t , is ( Mg g M short , ( , q t ) = 1 4 Dπ 0      2 ) t ( q − v + a 4 D 0 p e − − 2 q 4 D 0 + p e 0 + p e + 2 ) t ( q − v − a 4 D 0     . (S20) 02D and does not vary The variance of the distribution, with time, however, the interval between the peaks does gradually increase as time increases. At long times ( , at short times is equal to is written as t τ ), ), q t Mg ), q t ( ( Mg c 7 g M long , ( , q t ) = 2 q D eff − 4 e 1 4 D π eff 1 +     eq p D R 0 2 τ c a D t eff     1 − 2 q D eff 2       . (S21)   g M long , ( , q t ) converges to the Gaussian distribution with a variance of 2 effD . E. Dynamics of the multi-mode model For the multi-mode model, a given active matter is operated by a variable composed of iΓ state is located in the three discrete states: infinitesimal area dx , then the probability of finding the active matter can be written as ( ρ Γ . The PDF satisfies the conservation law, written as −Γ . If the active matter with a 0Γ , and +Γ , , i x t dx ) , ∞ ∑∫ dx −∞ 3 i ( ρ Γ i , , x t ) = 1 . (S22) From the conservation law, the continuity equation for the PDF is written as ∂ t ∂ ( ρ Γ i , , x t ) = − ( ( x  ρ Γ i , , x t ) ) + ∂ x ∂ 3 ∑ j ≠ i −Κ   ( ρ Γ i , , x t ) + Κ i → j ( ρ j → i Γ j , , x t )   , i j→Κ denote the time derivative of x and the rate constant from a state iΓ to where x and jΓ . Here, we consider the motion of active matter in an overdamped environment where acceleration is zero. The time derivative of an active matter position is written as dx dt ) = Γ + v ( 1 γ ξ− ⋅ ( ) t , (S23) ( )tξ represents the random fluctuating force modeled as Gaussian white noise. The over the Gaussian white noise gives the observed PDF [1]. The application of the cumulant expansion gives Eqs. (3) and (4), where we set ( ) v Γ , to for the sake of simplicity. Internal-state-dependent velocity, ( ρ Γ , , i x t ) ) ) where ensemble average of ( iP Γ ( ), ( iP x t iP Γ depends on the state as , x t , x t , , v ( ) +Γ = v a , ( v Γ = , and )0 0 v ( v −Γ = − . a ) F. Convergence of to ( , ) SP x t at the small R limit ( , ) MP x t ( R ≡ eq p 0 The population ratio, ( eq p + + eq p − ) = τ τ 0 a = k 0 2 k a ) , of the passive state to the active state modulates the shape of the probability density of the active matter, the multi-mode model. Applying written as to Eq. (S2) produces 1 k τ−= 0 c 1 1 R τ− − c and 2 a k = ( , ) MP x t , in ), (  MP w s , 8  , P w s M ( ) = = ( ( s + D 0 s + D 0 2 2 ( w ( s 2 w 1 − τ + c 2 w s + ( ) ( 2 D + 0 ( ) ( s w D 0 s D + 0 2 1 − τ + c 2 w D 0 w + R + 1 − τ c 1 1 − − τ c R + 1 − τ + c 2 w 2 )( ) 2 s w v w D + 0 a ) 1 1 1 − − − s D τ τ + + + 0 c c )( ) 2 1 1 − − v w Rs R w R D τ τ + + 0 c c a ) 2 1 1 1 − − − Rs R w R τ τ τ + + + + + c c c + )( + )( + D 0 + ( ) 1 R R + ) 2 2 1 1 − − v w v w R τ + a a c ( ) 2 2 1 R R + ) 2 2 2 1 − v w v R τ + a a c w 2 2 2 2 . (S24) In the limit of 0 R → , ),  MP w s ( is written as lim 0 R →  , P w s M ( ) = s + )( w s 2 ( s + D 0 w D 0 D + 0 2 + 2 w 1 − τ c 1 − + τ c ) + 2 2 v w a . (S25) The relaxation time of the single-mode model is produces ),  SP w s as ( τ c = ( 2 ak − ) 1 . Applying 1 2 a k τ−= c to Eq. (S10)  , P w s S ( ) = ( 2 2 + 0 2 D w s D w + )( s D w s + ),  MP w s + ( 0 0 is the same as in the small R limit. 1 − τ c 1 − τ + c ) + 2 2 v w a . (S26) ),  SP w s ( G. Relaxation time of the two solvable models In the high friction regime, where we can safely neglect the inertial term in the Langevin , of active matter with a friction constant, γ, can be written as the ( )x t equation, the velocity, sum of two components: 1 ( ) tγ ξ− ( ( )) t Γ ( ) x t  + = v s , (S27) ( ) sv Γ and 1 ( )tγ ξ− where represent the velocity component of a self-propelled, ballistic motion, which is dependent on the internal state, Γ, and the velocity component caused by the random fluctuating force. If we assume that the initial position of active matter is zero, then the time integration of Eq. (S27) produces the time-dependent position, written as ( ) x t = ( v s t ∫ 0 ( ( )) Γ τ γ ξτ τ ( ) + d 1 − . (S28) ) From Eq. (S28), the MSD of the active matter can be evaluated from the velocity correlation function, written as ) ) ⋅ v s ( Γ ( τ 1 ) ) + 1 2 γ ( ( ξτ ξτ 1 ) ⋅ 2 )    , (S29) ( τ 2 ( Γ v s    ) ( d τ τ − t 2 x ( ) t = t ∫ 0 d τ τ 1 d 2 t ∫ 0 t ∫ 0 = 2 D t 0 + 2 v s ( ) τ ⋅ v s ( ) 0 9 ) where we denote obtain the following equation: ( ) τΓ sv ( in short as ( ) sv τ . By comparing Eq. (S29) with Eq. (6), we ( d τ t ) − τ v s ( ) τ ⋅ v s ( ) 0 = ( τ − e D a c t τ c 1 − + t τ c ) . (S30) t ∫ 0 aD is equal to Because normalized time correlation function of velocity, 2 svτ c tφ , as sv ( ) , the second derivative of each side of Eq. (S30) provides the φ v s ( ) t = ( ) v t v s s (0) 2 v s = t − e τ c , (S31) where cτ is given by ( 2 ak − for the single-mode model and ) 1 1 0k − for the multi-mode model. H. General model In general, a given active matter moves in a multidimensional space, d, and its random pτ . To obtain analytic solutions for this general fluctuating force has a finite relaxation time, model, the velocity of active matter corresponding to Eq. (1) is generalized to r  ( ) t = v s ( ( )) t Γ + ξ 1 ( ) tγ− , (S32) where each bold symbol denotes the d-dimensional vector corresponding to each scalar quantity in Eq.(1). The integration of each side of Eq. (S32) from 0 to t produces the time- dependent position, , written as ( )tr r ( ) t = ( v s t ∫ 0 ( ( )) Γ τ γ τ τ ( ) ξ 1 − + ) d , (S33) where we assume the initial position is zero. From Eq. (S33), the MSD is written as 2 r ( ) t = = = t t 2 ∫ 0 ∫ 2 d d ( ) t r ξ 0 ( 2 − d τ τ γ − ) t (0) ⋅ ξ ξ ( ) τ   ) ( ) τ τ φ τ − 1 − p ξ ( τ t  D  0 ( ) t s 2 2 + r v + v s ( ) τ ⋅ v s (0 )   + D 1 ( ) − τ φ τ a c v s   , (S34) where ( )tφξ denotes the normalized time correlation function, random fluctuating force, ( )tξ , and the relaxation time, , of the 2 ξ ξ ξ ( ) t ⋅ (0) pτ , is defined as τ p γ ∞ ∫ 2 − is (0) φ∞≡ ∫ dt 0 ξ ξ ( ) (0) dt t ⋅ defined 1 − 0 ξ ( ) t . , Here, the diffusion coefficient for passive motion is defined by by and the ∞ ∫ 1 −= D d . The MSD consists of two independent movements from the a diffusive mode and the self-active mode. The diffusive mode contribution to the MSD is D d 0 self-propelled motion diffusion v v dt coefficient for (0) ( ) t = ⋅ 0 s s 10 2 ( ) t ( ) t 2 t 0 ∫ 1 − τ a c ξ ) defined as r ξ ≡ 2 1 − dD τ 0 p ( ( ) d τ τ φ τ − ) t , while the self-propelled mode contribution is ≡ r v defined as ( )tξ is Gaussian, then the analytic solutions for the fourth moment of displacement can be written as . If we assume that the distribution of ( ) d τ τ φ τ dD ∫ 2 − t v 0 s s t ( 4 r ( ) t = ( 1 2 + d 1 − ) 2 r ξ ( ) t 2 + ( 2 1 2 + d 1 − ) 2 r ξ ( ) t 2 ( ) t r v s + 4 ( ) t r v s , (S35) v s ( t 4 ) ⋅ v s ( ) t 3 v s ( t 2 ) ⋅ v s ( ) t 1 . The non-Gaussian 4 t t ( ) t s ≡ vr 4! where dt 1 parameter for the general model is written as 2 dt dt dt ∫ ∫ ∫ ∫ 2 0 4 0 0 3 0 4 2 t 3 t α R ( ) t ≡ d + 2 d with 4 ( ) r t ( ) t r 2 1 − = 2 α ≡ ( ) t v s d + 2 d 2 ( ) t r v s 2 r ( ) t 4 ( ) t r v s 2 ( ) t r v s α v s 2 ( ) t − 1 2 . (S36) R tα approaches zero because ( ) 2 r ( ) t 2 r v s 2 ( ) t 2 . At long times, At short times, R tα also approaches zero because ( ) ( ) tαv s approaches zero. I. Time correlation functions for two solvable models Time correlation functions of the velocity component, sv , of self-propelled motion are ( )tα as well as the second and fourth moments of displacement. sv is only dependent on the internal state, we analytically obtain the aD and used to calculate In our model, because time evolution of internal state probabilities as ( ) P t + ( ) P t −         = G     P + P − ( ) 0 ( ) 0     with G = E − k t a cosh sinh ( k t a ( k t a ) )    sinh cosh ( ( k t a k t a ) )    (S37) for the single-mode model and ( ) P t + ( ) P t 0 ( ) P t −           = G P + P 0 P −      ( ) 0 ( ) 0 ( ) 0      with 11 G = 1 2 + k a k 0         − ( k 0 + 2 k a ) t k 0 e k a + k k 0 0 + − 2 k a k 0 ) t − ( k 0 e k a + + e − + 2 k a ) e − k t 0 k a 0 ( k 2 k − − 0 2 k a + 2 k a 0 − k t 0 ) e ( k 2 − k − e ( k 0 + 2 k a ) t k a + a 2 e k a k 0 + − ( k 0 + 2 k a ) t k a − k e a − ( k 0 + 2 k a ) t k a + + 2 k a ) e − k t 0 0 ( k 2 k − − 0 2 k a − ( k 0 + 2 k a ) t k 0 e k k 0 0 + − 2 k a k 0 ) t − ( k 0 e − e + + 2 k a 0 − k t 0 ) e ( k 2         (S38) for the multi-mode model. We obtain the velocity autocorrelation function, through the following equation: ) ( ) v t v s s ∑ ∑ (0) ,0 ,0 G v s v s P Γ Γ Γ Γ Γ = ( ) ( ) ) ( ( t , i i j i j , (S39) v s ( ) t ⋅ v s (0) , , i ∈+ − ∈+ − j , ) ( i j Γ G ,0 iΓ to , tΓ denotes a transition matrix from jΓ after t time passing. The where transition matrices are shown in Eq. (S37) for the single-mode model and in Eq. (S38) for the multi-mode model. The calculation results of for the single- mode model and for the multi-mode model, which coincides with Eq. (S31). By calculating the MSD through the application of the calculated velocity autocorrelation function to Eq. (S29), we obtain Eq. (6). The four-time velocity autocorrelation function, ( ) v t v t v t v t 1 s , is obtained by the following equation, written as 2 a ap v e− ( ) v t v s s 2 2 av e− are ( ) 3 (0) ak t k t 0 ( ( ) ) 2 4 s s s s 4 ( ( ( ) 3 ) ( ) ) v t v t v t v t 1 s ∑ ∑ ∑ ∑ ( v = s 2 s s , j , , i ∈+ − ∈+ − ∈+ − ∈+ − G , t × l Γ , Γ k t , ( 4 k l 3 ) G Γ ( ) v s ( Γ k ) v s l Γ , t 3 Γ , t 2 j k j ( Γ ) G ) ( v s ( Γ i ) . Γ , t 2 j Γ i , ) t P 1 ( Γ i , t 1 ) The calculation results of ) model and + ( k t 3 0 2 4 p v a a e − t 2 ( ( ( ) v t v t v t v t for the single-mode 1 s for the multi-mode model. The four-time e R are 2 ( k t 0 2 av e− 4 t − + − 3 ) ( 4 k ( ) s ) ak 2 − t 1 t 3 t 1 − − − s s t t 2 4 t 4 a ( ) 3 ) ) e 2 ( ) velocity autocorrelation functions can be used to generate ( ) 4 t vr s in Eq. (S35), and their results are equal to the fourth moment of displacement which are written in Eqs. (S7) and (S12). J. Derivation of short time mean velocity distribution from Eq. (1) In our model, the velocity of active matter consists of two components in Eq. (1). If we assume that the two components are independent, then the mean velocity distribution is written as ( , f v t ) = ∫ ( dv dv δ ξ s v − ( v s + v ξ ) ) f s ( , v t s ) , f v t ξ ξ ( ) , (S40) 12 sv and f ( ), v t s s respectively denote the velocity component caused by self- where propelled motion and its distribution function; vξ and denote the velocity component due to the random fluctuating force and its distribution function. At short times, the velocity component, , caused by the random fluctuating force is already relaxed and 02D t , whereas the self-propelled motion follows a Gaussian distribution with a variance of approximately maintains its direction. The two distribution functions at short times can be written as ), v t f ξ ξ 1 ( )tγ ξ− ( ( , f v t ξ ξ ) = e − 2 v ξ 04 D t 4 π D t 0 and f s ( , v t s ) = ( eq p δ i v s − v i ) ∑ i ∈Γ . (S41) By applying Eq. (S41) to Eq. (S40), the mean velocity distribution function at short times ∞ −∞ ∫ du dv dv e ξ s ∫ ( iu v ( − v s + v ξ ) ) ( eq p δ i v s − ∑ i ∈Γ ) v f i ( , v t ξ ξ ) can be rewritten as 1 2 π , v t short = ) ( f eq p i 1 = ∑ ∫ 2 π ∑ eq p i i ∈Γ = ∫ i ∈Γ ∞ −∞ ( iu v v − i ) due − iuv ξ dv e ξ ∫ ( , f v t ξ ξ ) dv f , v t ξ ξ ξ ( ( ) δ v − − v i v ξ ) = ∑ i ∈Γ eq p f i ξ ( v − , v t i ) = ∑ i ∈Γ eq p i 1 4 D tπ 0 − ( )2 v v − i 04 D t e . (S42) Eq. (S42) is equivalent to the mean velocity distribution for the multi-mode model at short times, which is shown in Eq. (S18). K. Stochastic simulation method Our stochastic simulation method consists of both the Brownian dynamics for the time evolution of an active matter position and the Gillespie method for the stochastic transition between internal states [2,3]. For the Brownian dynamics, we numerically integrate Eq. (S23) as ( x t ) + ∆ = t ( ) x t + Γ ⋅∆ + v t 02 D t ⋅∆ ⋅ ( ) tξ ' , (S43) ( ) x t , t∆ , and ' tξ denote the active matter position at time t, the size of the time step, where , respectively [3]. For the Gillespie method, we and the Gaussian random number with assume that the transitions between internal states for the multi-mode model are forbidden, except the following four unimolecular reactions: transitions described by ( )0,1G those ( ) ( ) 13 + →Κ = k − →Κ = k 0 0 +Γ →Γ , Γ →Γ [2]. The 0 0 reaction constants for the forbidden transitions are set equal to zero. Our stochastic simulations proceed as follows: Γ →Γ , and 0 −Γ →Γ , − + 0 0 0 ak→ −Κ = 0 ak→+Κ = 0 1. Randomly choose an internal state of active matter based on the equilibrium population between states and set the initial position equal to zero. Set the selected state to the current state, cΓ . 2. Based on the current state, calculate the waiting time for a reaction using the equation, , where RN denotes an evenly distributed random number RN = − τ ln ) ( Κ∑ j c ≠ c → j between 0 and 1, because concentration of the selected state is 1 and the concentration 0Γ state has two reaction paths with equal for the other states is zero. Only the probability, and the other states have only one path for state transition. 3. Until the waiting time τ is over, evolve the time-dependent position using Eq. (S43) with the state-dependent velocity ( v Γ and a given time interval )c t∆ . 4. After finishing the time evolution in step 3, change the current state to the state determined by the transition in step 2. Return to step 2 when the elapsed time of the trajectory is less than the time limit of the trajectory. 5. Return to step 1 until sufficient trajectories are collected. We use 500,000 trajectories to obtain the velocity distributions and the second and fourth moments of the displacement distributions. L. References H. Risken, The Fokker-Planck equation : methods of solution and applications [1] (Springer-Verlag, New York, 1996), 2nd edn., Springer series in synergetics,, 18. [2] [3] D. T. Gillespie, J. Phys. Chem. 81, 2340 (1977). D. L. Ermak, J. Chem. Phys. 62, 4189 (1975). 14 ( ) 0 a ( ) c ( ) M tα in all ranges of R and t τ are plotted when Figure S1. Non-Gaussian parameter. D D is equal to 0.1. The distribution of displacement is sub-Gaussian (super-Gaussian) in 0 all time ranges if R<1.62 (R>2.00), as shown in Fig. 3(d). The two horizontal red lines represent M tα for the two critical values of R: 1.62 and 2. In the range 1.62 < R < 2.00, the displacement distribution switches from sub-Gaussian to super-Gaussian along the time axis. The boundary between the sub-Gaussian and the super-Gaussian distribution is represented by the black line. 15
1507.06182
3
1507
2016-05-04T09:57:27
Role of turn-over in active stress generation in a filament network
[ "physics.bio-ph" ]
We study the effect of turnover of cross linkers, motors and filaments on the generation of a contractile stress in a network of filaments connected by passive crosslinkers and subjected to the forces exerted by molecular motors. We perform numerical simulations where filaments are treated as rigid rods and molecular motors move fast compared to the timescale of exchange of crosslinkers. We show that molecular motors create a contractile stress above a critical number of crosslinkers. When passive crosslinkers are allowed to turn over, the stress exerted by the network vanishes, due to the formation of clusters. When both filaments and passive crosslinkers turn over, clustering is prevented and the network reaches a dynamic contractile steady-state. A maximum stress is reached for an optimum ratio of the filament and crosslinker turnover rates.
physics.bio-ph
physics
Role of turn-over in active stress generation in a filament network Tetsuya Hiraiwa1,2,3,Guillaume Salbreux1,4 1 Max Planck Institute for the Physics of Complex Systems, Dresden, 01187, Germany 2 Fachbereich Physik, Freie Universitat Berlin, Berlin, 14195, Germany 3 Department of Physics, Graduate School of Science, The University of Tokyo, Tokyo, 113-0033, Japan and 4 The Francis Crick Institute, 44 Lincolns Inn Fields, London, WC2A 3LY, United Kingdom (Dated: August 9, 2021) We study the effect of turnover of cross linkers, motors and filaments on the generation of a contractile stress in a network of filaments connected by passive crosslinkers and subjected to the forces exerted by molecular motors. We perform numerical simulations where filaments are treated as rigid rods and molecular motors move fast compared to the timescale of exchange of crosslinkers. We show that molecular motors create a contractile stress above a critical number of crosslinkers. When passive crosslinkers are allowed to turn over, the stress exerted by the network vanishes, due to the formation of clusters. When both filaments and passive crosslinkers turn over, clustering is prevented and the network reaches a dynamic contractile steady-state. A maximum stress is reached for an optimum ratio of the filament and crosslinker turnover rates. 6 1 0 2 y a M 4 ] h p - o i b . s c i s y h p [ 3 v 2 8 1 6 0 . 7 0 5 1 : v i X r a 2 The cell cortical cytoskeleton is essential in processes involving cell shape changes [1, 2]. In the cortex, myosin molecular motors are assembled in bipolar filamentous structure which bind to actin filaments and generate forces by consuming the chemical energy of the hydrolysis of adenosine triphosphate (ATP). The action of myosin motors result in the generation of an active, contractile stress, whose spatial distribution in the cortex plays a key role in cellular morphogenetic processes [3, 4]. In living cells, passive, active crosslinkers and actin filaments are continuously exchanged between the cortex and the cytosol [1]. As a result, cytoskeletal networks can release elastic stresses stored in the network and undergo large-scale flows. Significant progress has been achieved trough in vitro studies and theoretical analysis of actomyosin networks to understand stress generation in networks with permanent filaments and fixed or unbinding crosslinkers [5 -- 11]. It is unclear however what is the role of turnover in stress generation and how filament networks can simultaneously rearrange and exert a permanent internal active stress. In vivo experiments suggest that the rate of turn-over is a major determinant of force generation by actomyosin networks [12, 13]. We ask here how the rate of turn-over of passive crosslinkers and actin filaments influence the active stress generated by motors in the network. We study a simplified mechanical model for a cytoskeletal network in two dimensions whose constituents are turning over (Fig. 1a). Filaments (actin filaments) and motors (myosin motors) are treated as rigid rods. Filaments are assumed to have a polarity represented by the arrows as shown in Fig. 1. Passive crosslinkers are assumed to be point-like and constrain the position of the filaments on which they are attached. Filaments are able to rotate freely around the cross linker position and around motor heads. Motor heads exert an active force fm with a constant magnitude fm = f0 on filaments, oriented toward the reverse direction of the arrow (the minus end of actin filaments). FIG. 1. (a) Schematic illustration of a 2D cytoskeletal network (left). Filaments are represented by black arrows, motors by red bars and cross linkers by green dots. The network exerts a stress σ on the boundaries of the box, arising from tensions acting within the filaments (ff ) and within the motors (fm). Motors move towards the arrowhead of filaments (for actin, the arrowhead corresponds to the barbed end). Network components stochastically exchange with a reservoir (right). (b) Motors move on filaments and rearrange the network on a timescale τ . Crosslinkers and actin filaments turn over on longer timescales τc > τ and τf > τ . To obtain forces acting on filaments, we introduce the effective mechanical potential, U = W +λ·g(xf,i, nf,i, xm,k, nm,k), where W = −f0Phk,ii ski is the work due to motor active forces, with ski the position of the k-th motor head on the i-th filament relative to the filament centre of mass, and the sum Phk,ii is performed for all the pairs of filaments (i) and motors (k) connected with each other. In addition, geometrical constraints arise from the conditions that cross linkers and motor heads are firmly attached to filaments, and that motor filaments have a fixed length. The coefficients λ are Lagrange multipliers imposing these constraints, g(xf,i, nf,i, xm,k, nm,k) = 0, where xf,i and nf,i (resp. xm,k, nm,k) indicate the centre of mass position and orientation unit vector of the i-th filament (resp. k-th motor). The motion of motors is taken into account by writing that the position of attachment of the k-th motor ski relative to the centre of mass of the filament i follows the dynamic equation µ dski dt ∂U ∂ski = − (1) 3 with µ a scalar friction coefficient arising from translational friction between the motor heads and the filament [14 -- 16]. Motors have a typical velocity vm = f0/µ, and we introduce a reference timescale τ = lf µ/(2f0). We neglect viscous forces arising from the fluid around the network compared to motor-filament friction, and the position and orientation of filaments xf,i, nf,i are relaxed instantaneously. To fix ideas, cortical actomyosin networks in a cell have a mesh size ξ ∼ 20 − 250nm [17, 18] and the typical cell diameter is several tens of micrometers. We therefore expect actin filaments to have a length lf of order 0.1 − 1µm, smaller than their persistence length lp = 16µm [19]. Myosins move on actin filaments with velocity vm ∼ 0.1− 3µm/s [20, 21]. The characteristic time for the myosins to move on a filament is τ ∼ lf /vm ∼ 0.03 − 10s. To compare this timescale to the effect of viscous stresses arising in the solvent of viscosity η, we note that the velocity of a filament in the network subjected to a force f0 is ∼ f0 ln(ξ/rf )/(ηlf ), giving a timescale τv ∼ l2 f η/(f0 ln(ξ/rf )), with rf ≃ 5nm the radius of a filament. Taking the force exerted by a myosin f0 ∼ 6 − 12pN , η ≃ 10−3P a.s for water, we find τv ∼ 10−4s ≪ τ . Finally, we expect the characteristic times τc and τf of the crosslinker and actin filament turnover, respectively, to be of the order of 10 s- 1 min [22, 23], slow compared to these timescales, such that τc ≫ τ and τf ≫ τ . a. Configuration with one motor and two filaments. We start by discussing the dynamics of configurations in- volving one motor attached to two filaments (Fig. 2a). The two filaments can be connected to a crosslinker, itself connected to the external network, which we consider here to be fixed in space. We distinguish several basic possible configurations, depending on the number and positions of attached passive crosslinkers (Fig. 2b): both filaments can be completely free (i), the two filaments can be crosslinked to each other (ii), or they can be crosslinked to the external network (iii)-(v). To investigate how molecular motors modify the filaments organisation, we study the relaxation to final state of these different configurations. By averaging over possible initial configurations, we evaluate whether motors form on average positive or negative force dipoles in the network (Fig. 2c). The motor force dipole is d = −(lm/2)n· (f1 − f2), with n the unit vector giving the motor orientation, and f1 and f2 are the forces by which the motors pull or push the filaments (Fig. 2d). In case (v), f1 and f2 include not only the forces exerted by the motor themselves but also the forces originating from the geometrical constraint (Appendix A). We find that possible initial configurations can be classified in 2 categories (Fig. 2b). In cases (i) and (iv), filaments are moved relative to each other until the motor detaches, so that a motor-induced force dipole acts on a transient time τ . This force dipole is expansile on average. In the quasi-static limit where τ ≪ τc, the contribution of transient filament-motor configuration to the overall stress is negligible compared to steady motor configurations. When filaments either (ii) have only one fixed attachment point to the external network, (iii) are connected to each other by a cross linker, or (v) have both more than 2 cross linkers, they relax to a steady configuration where the motor exerts a constant force dipole. By averaging the resulting force dipole over possible initial configurations of the two filaments and the motor (Appendix A), we find that the motor exerts a contractile force dipole on average in cases (iii) and (v) (Fig. 2c). As in Ref. [7], the bias towards contractile states arises from instabilities of expansile configurations (Appendix A). The average force dipole hdi vanishes for point-like motors, lm → 0 [24]. b. Numerical simulations of networks with turnover. We next numerically simulated a network of Nf filaments, Nm motor rods and Nc passive crosslinkers in a square box of width W with periodic boundaries (Fig. 1). The frame of the box is not allowed to deform. We fixed the normalised motor density l2 mc to 1, where c = Nm/W 2 is the motor density. To make numerical simulations easier, the geometrical constraints g were replaced by linear springs mimicking the contacts at the junctions between motors and filaments, and at crosslinking points. Initial conditions are obtained by randomly positioning filaments in the network, and timescales are normalized to the reference time τ . Turnover is introduced by stochastically removing cross linkers, motors and filaments from the network with rates 1/τc, 1/τm and 1/τf . For simplicity, turnover rates of crosslinkers are taken here independent of the forces they sustain [11]. Filaments, motors and passive crosslinkers are added in the network with on-rates kf on. New filaments take random positions and random orientation, motors take a randomly chosen position on two filaments points separated by a distance lm, and passive crosslinkers are put on a randomly chosen filament intersection. on and kc on, km We evaluate the components of the stress tensor σij (i, j = x or y) acting on the boundary of the box (Fig. 1a). ij obtained by summing forces acting both ij from forces acting within motor rods crossing the boundary of the box (Fig. 1). In a ij = 0 in the absence of large-scale deformation. In a non-linearly elastic material [25] The total stress is given by two contributions σij = σf within the filaments and σm linear elastic or viscous material, σf however, σf and Appedix B) ij 6= 0 (Appendix B). At steady state, the average stress acting within the motors is given by (Ref. ij + σm ij , with σf where n, d and c denote the orientation, motor-induced force dipole strength and concentration of bound motors. When the dipoles are isotropically oriented, σm ij = σmδij with σm ≃ hdci/2. σm ij ≃ hdninjci , (2) 4 ✄✂ f0 f0 1st actin filament 2nd actin filament ☎✞✆✝ f0 f0 f0 f0 f0 f0 f0 f 0 f0 f0 ✘✙✚ ✛✜✜✜✢ ✛✜✜✢ 10 20 30 2 1 0 -1 0 ✠✂ ✡☛☞✌✍✎✏✌✑✒✓ ✔✕✖✎☞✗✑✒✓ ✁✂ ☎✞✝ ☎✞✞✝ ☎✞✞✞✝ ☎✆✝ ✟✂ FIG. 2. Contractile and expansile configurations for one motor and two filaments. (a) Two filaments are attached by fixed cross linkers to the external network. (b) (Left) Possible initial configurations for one motor, two filaments and one or several cross linkers, and final configurations. (c) Average strength of the contractile force dipole generated by the molecular motor hdi in the final configurations (ii), (iii), (v) where the motor does not detach, as a function of the filament length lf . (d) Contractile (force dipole d > 0) and expansile (d < 0) configurations. We first performed simulations without turnover of filaments. We focus on the isotropic component of the stress σ ≡ (σxx + σyy)/2. Figure 3a (blue curve) shows the typical time evolution of the stress, for permanent crosslinkers and for crosslinkers with a finite lifetime. Without crosslinker turnover, the system reaches a stationary state where the stress σ fluctuates around a finite positive value. Figure 3b shows the average value of the resulting stress (circles) as a function of the number of cross linkers Nc. For a small number of crosslinkers (Nc → 0), no stress is observed, σ = 0. Above a critical value of the number of crosslinkers Nc > N ∗ c ≃ Nf , a transition occurs and c but Nc < N ∗ a positive contractile stress appears in the system. When Nc ∼ N ∗ c , although a portion of motors generate contractile force dipoles, filaments aggregate in clusters and hence σ ∼ 0, suggesting that the transition is 5 associated to the network connectivity (Supp movies M1 and M2). The stress then further increases with the number of crosslinkers and eventually saturates to a positive value. Such a transition to contractility as a function of the number of crosslinkers has been reported in in vitro reconstituted networks [10, 26], as well as network clustering [27]. The average stress only within the motors σm is also plotted in Fig. 3b (cross marks). A similar transition occurs for a critical value of the number of cross linkers. The stress σm is however larger than the total stress for a large number of cross linker Nc > Nf , implying that the filament network is under compression (σf < 0). For large Nc, configurations (v) dominate in the network (Fig. 3c). The saturating value of σm/σ0 ∼ 0.6 is nevertheless smaller than the average force dipole obtained by averaging all possible filament orientations hd/2i/(f0lm) ∼ 1.0 (Fig. 2c); this is because some configurations do not relax to equilibrium on the characteristic timescale of myosin turnover. When crosslinkers are allowed to turn over, the average stress first reaches a positive value before decaying to zero (Fig. 3a, red and green curves), even though the stress within the motors σm is still non zero (Fig. 3d). The decay of the stress correlates with the collapse of the network in an isolated cluster, where filaments accumulate (Fig. 3e, Supp movie M3). To evaluate the decay timescale τst, we fitted an exponentially decreasing function σ(t) = A exp(−t/τst) to the simulation results with fitting parameters A and τst. The decay timescale of the stress increases exponentially with Nc for large Nc (Fig. 3f). The relaxation timescale of the stress can be understood as follows: the timescale τst corresponds to the relaxation Maxwell time on which the network becomes fluid, due to turnover of passive crosslinkers enabling network rearrangements. This time can be estimated by τst = τc n∗ c (en∗ c − 1) (3) with n∗ c = 2Nc/Nf . To obtain Eq. (3), we assume that filaments with at least one crosslinker are fixed, and only filaments with no crosslinker attachment can rearrange. We then compute the first passage time at which a filament is free from cross linkers, starting from a configuration where the filament is attached with only one crosslinker (Appendix C). Equation (3) accounts for the characteristic time of stress decay in the network for large values of Nc/Nf (Fig. 3f). We now turn to simulations where both filaments and crosslinkers turn over. In the cell, actin filaments polymerise and depolymerise. Here, we account for this process by simply introducing a rate of filament turnover, τ −1 . Passive crosslinkers and motor heads are removed together with the filaments to which they are attached. Remarkably, with both crosslinker and filament turnover, the network evolves towards a steady state with a non-zero positive stress (Fig. 4a-b). As in the previous case, a transition from a non-contractile to a contractile network appears when the number of crosslinkers Nc is increased (Fig. 4c, Supp movies M4 and M5). Note that the stress σ deviates again from the average stress within the motors σm (Fig. 4c, inset). The dependency of the fraction of configurations as a function of Nc is qualitatively similar to the case without filament turnover (Fig. 4d). f We then varied the filament turnover timescale τf (Fig. 4c,e,f). We find that the stress reaches a maximum for intermediate values of the filament turnover timescale (broken line in Fig. 4e, and Fig. 4f): for slow filament turnover τf ≫ τst, the network collapses in clusters due to crosslinker turnover, while for fast filament turnover τf ≪ τst, filaments are removed before motors reach a configuration where they can exert a force dipole in the network. We find that the optimum value of the ratio of turnover time scale in Fig. 4f, τf /τc ∼ 10, is of the order of the ratio experimentally measured values of turn-over of actin and crosslinker in the cell cortex, τf ∼ 15 − 45s and τc ∼ 7 − 14s [1]. c. Conclusion. We propose the following dynamic picture for the stress generated in a rearranging network: molecular motors move on filaments on a timescale τ , fast compared to the crosslinker turnover τc. When molecular motors bind to pairs of filaments, they either displace them and detach, or find steady configurations where they generate predominantly contractile force dipoles (Figs. 2 and 3a-c). In networks with permanent crosslinkers, a transition to contractile state of the network occurs for a large enough number of crosslinkers (Fig. 3b). When crosslinkers are allowed to turnover, the network can rearrange and flow, filaments collapse in a cluster, and the total stress in the network vanishes even though the stress only within the motors is still contractile (Fig. 3a,d-f). Introducing filament turnover occurring on a timescale τf comparable to the crosslinker turnover timescale τc prevents this clustering mechanism, and allows the network to generate a steady-state contractile stress (Fig. 4). It will be interesting to investigate stress generation in in vitro experiments where crosslinkers and filaments are allowed to turn over. Acknowledgements: We thank Matthew Smith for helpful discussions. This work was supported partly by the JSPS Institutional Program for Young Researcher Overseas Visits (T.H.), the Postdoctoral Research Fellowship of the Alexander von Humboldt foundation (T.H.), the JSPS Core-to-Core Program (T.H.), the Max Planck Gesellschaft (G.S.), and the Francis Crick Institute which receives its core funding from Cancer Research UK, the UK Medical Research Council, and the Wellcome Trust (G.S). 6 (a) Time evolution of the FIG. 3. Stress in a network of motors and filaments, with and without crosslinker turnover. normalised stress σ/σ0 for τ −1 c = 0, Nc = 1.2Nf (blue), τc = 100τ , Nc = 1.6Nf (red) and τc = 100τ , Nc = 1.2Nf (green). (σ0 = f0lmNm/W 2) (b) Steady-state isotropic stress σ (circles) and motor stress σm (cross marks) as a function of the crosslinkers number Nc for τ −1 c = 0. (c) Fraction of different configurations (i)-(v) in Fig. 2 at steady state as a function of the crosslinker number Nc for τ −1 c = 0. (d) Isotropic stress σ as a function of the crosslinker number Nc, for finite crosslinker lifetime τc = 100τ , and for several simulation times (τsim. = 300, 600, 1, 200 and 2, 400τ ). Stress within the motors σm is shown in the inset for τsim. = 300 and 2, 400τ . (e) Snapshots of a simulated network with crosslinker turnover (Nc = 1.2Nf , τc = 100τ ). (f) Stress decay time τst as a function of the crosslinker number Nc with crosslinker turnover (τc = 100τ ). Solid line, theoretical prediction (see main text). Other parameters: Nf = 1, 000, Nm = 100, τ −1 onτc = 20, lf = 2lm, W = 10lm, τm = 100τ and km f = 0, kc onτ = 20. Appendix A: Calculation of the average force dipole for two filaments We detail here how we obtain the average force dipole exerted by a motor on two filaments in the configurations shown in Fig. 2. As mentioned in the text, we assume here that crosslinkers are rigidly fixed to the external network. The geometry of the two filaments and motor can be described the position of the motors on the two filaments s1, s2 and by the two angles ψi between the motor filament and i-th filament (i = 1, 2). The motor positions si evolve ✆✂ 7 ✁✂ 0.5 0.4 0.3 0.2 0.1 0 -0.1 0 200 400 600 800 1000 1200 t=400τ t=1200τ . 8 0 4 0 . . 0 0 0.0 1.0 2.0 ✝✂ 0.5 0.3 0.1 −0.1 ✞✂ n o i t c a r F s f i t o m f o 1.0 0.8 0.6 0.4 0.2 0.0 ✟☛✡ ✟✠✡ ✟☛☛☛✡ ✟☛✠✡ 0.0 1.0 2.0 3.0 0.0 1.0 2.0 ✟☛☛✡ 3.0 100 ✄✂ 10 1 0.1 0.01 0.0 0.3 0.2 0.1 0.0 ☎ ✂ 0.04 0.0 2.0 3.0 0.01 1 100 FIG. 4. Stress in a network with filament turnover. (a) Time evolution of the stress in a network with crosslinkers and filament turning-over for τf = 100τ , τc = 100τ and Nc/Nf = 1.6. (b) Snapshots of the network evolution corresponding to (a). (c) Steady-state stress σ as a function of the number of crosslinkers for τc = τf = 100τ (circles), τc = τf = 10τ (triangles) and τc = τf = 1τ (squares). Inset: steady-state stress σ (circles) and motor stress σm (cross marks) for τc = τf = 100τ . (d) Fraction of different configurations (i)-(v) at steady state, as a function of Nc (τc = 100τ , τf = 100τ ). (e) Heat map for contractile stress σ as a function of Nc and τc/τf (τc = τ ). The broken line indicates τf = τst. (f) Stress as a function of the ratio of filament and crosslinker turnover rate (τc = τ , Nc = 1.6Nf ). Other parameters as in Fig. 3, except τ −1 m = 0. according to while the angles ψi relax instantaneously, so that ∂U/∂ψi = 0. µ dsi dt ∂U ∂si , = − (A1) 1. No crosslinkers attached, or one filament is attached by only one crosslinker and the other filament is rigidly attached to the external network Case (i) and (iv): In these situations, no steady-state is reached as the motor always unbinds from the two filaments. 8 A case (ii): B Parallel filaments: C Anti-parallel filaments: [1] EXPANSILE, UNSTABLE [2] EXPANSILE, STABLE [3] CONTRACTILE, STABLE [4] CONTRACTILE, UNSTABLE [3] [1] [4] [2] [3] [1] D case (iii): E Parallel filaments: F l >am u n s t a b l e [6] s t a b l e Anti-parallel filaments: u n s t a b l e u n s t a b l e [2] u n [7] s t a b l e u n s t a [5] [1] EXPANSILE, UNSTABLE [2] EXPANSILE, STABLE s t a b l e u n s t a [3] b l e b l e s t a b l e [6] u n s t a b l e u n s t a b l e Contractile Contractile ✎✏✑✒ u n s t a b l e [3] EXPANSILE, UNSTABLE [4] CONTRACTILE, STABLE [5] CONTRACTILE, UNSTABLE l <am u n s t a b l e u n s t a b l e u n s t a b l e ✎✏✓✒ Expansile ✎✏✑✒ Expansile ✎✏✓✒ [6] s t a b l e [7] u n s t a b l e u n s t a [1] [4] s t a b l e u n s t a [3] b l e b l e s t a b l e [6] u n s t a [6] CONTRACTILE, STABLE [7] CONTRACTILE, UNSTABLE b l e u n s t a b l e u n s t a b l e Contractile ✎✏✑✒ Expansile G case (v): H I [1] ✁✂✄☎✆✄✝✞✟✠ ✡✄✆☛✞✟ [2] ✟☞✌✆✂✡✝✞✟✠ ✍✂✡✄✆☛✞✟ [1] ✎✏✑✒ Contractile Expansile ✎✏✓✒ ✎✏✓✒ [2] FIG. 5. Final configurations reached by two filaments bound by a motor, with the filaments attached to a rigid external network with a varying number of cross linkers. A, D, G, example configurations for each case, according to the number of cross linkers; B, E, H, corresponding possible steady-state configurations; C,F,I, Phase plot of the dynamics of the motor according to the coordinates of its position on the two filaments, x1, x2. Red lines and dots correspond to stable steady-state configurations, while dotted black lines correspond to unstable steady-state configurations. Green regions represent values of the coordinates x1 and x2 which are not geometrically accessible. 2. Two filaments attached to each other by a crosslinker Case (ii): When the two filaments are attached by a single crosslinker, the mechanical potential U reads U = −f0(x1 + x2) + λ(cid:18)qx2 1 + x2 2 − 2x1x2 cos θ − lm(cid:19) (A2) with x1 and x2 the distances between the myosin attachment points to the crossing point of the two filaments, taken positive in the direction of the filament towards which motors move (Fig. 5A). θ is the angle between the two filaments, such that n1.n2 = cos θ, with ni the unit vector giving the orientation of the filament i. In addition, λ is a Lagrange multiplier ensuring that the length lm of the motor is fixed. Three situations can then occur: • when θ 6= 0 and θ 6= π and the filaments are not parallel neither antiparallel, the angle θ is given by cos θ = 1 + x2 x2 2 − l2 m 2x1x2 (A3) and is free to adjust as the motor position changes. The evolution of the position of the motor on the filaments is given by µ µ dx1 dt dx2 dt = f0 = f0 9 (A4) (A5) such that the motor moves until it detaches from the filaments, or until the filaments become parallel or antiparallel. • When the two filaments are parallel and point in the same direction, θ = 0 and (x1 − x2)2 = l2 mass of the motor at position (x1 + x2)/2 follows the dynamic equation m. The center of µ d dt x1 + x2 2 = f0 (A6) and the motor moves on the filament until it detaches. • When the two filaments are antiparallel, θ = π and (x1 + x2)2 = l2 m. Two configurations are then possible, according to whether the two filaments point away or towards the center of the motor from their attachment point to the motor. These two configurations correspond respectively to x1 + x2 = −lm (expansile configuration) and x1 + x2 = lm (contractile configuration). Any position of the motor on the filament is then a steady-state solution, as long as the two motor ends each bind on the filaments. To test for the stability of these two configurations, we consider a slight change of the angle between the two filaments, θ = π + δθ, with δθ ≪ 1. We take the initial position of the motor to be x1 = x0 + χlm/2 and x2 = −x0 + χlm/2, with χ = −1 for an expansile configuration and χ = 1 for a contractile configuration. x0 = (x1 − x2)/2 is the position of the center of mass of the motor, relative to the crosslinker joining the two filaments, measured positively in the direction of the first filament. We find then the following dynamic equation for δθ: µ dδθ dt = dx1 dt = f0 + ∂θ ∂x1 8χlm m − 4x2 l2 0 ∂θ ∂x2 dx2 dt 1 δθ (A7) where we have used Eqs. (A3), (A4) and (A5) from the first to the second line. The sign of the right hand side of Eq. (A7) indicates whether the angle θ increases or decreases when the two filaments are slightly rotated away from their antiparallel configuration. Therefore, the associated configuration is unstable when the sign is positive, and stable otherwise. We find therefore that the stability of the configuration depends on the position of the center of mass of the motor: for x0 < lm/2 (motor near the crosslinker), the expansile configuration is stable, and the contractile configuration unstable. For x0 > lm/2 (motor away from the cross-linker), the expansile configuration is unstable and the contractile configuration is stable. The flow diagram in Fig. 5C case (ii) shows the corresponding dynamics in the space of the motor position (x1, x2). Red segments indicate the stable steady states. In the green regions, no value of the angle θ allows for the motor to have position (x1, x2) on the two filaments. 3. Two filaments, each attached by a crosslinker to the external network The mechanical potential reads in that case U = −f0(x1 + x2) + λ(cid:20)qx2 1 + x2 2 + l2 m + 2x1lm cos ψ1 + 2x2lm cos ψ2 + 2x1x2 cos(ψ1 + ψ2) − a(cid:21) , (A8) with x1 and x2 the distances between the myosin attachment point and the crosslinker position on each filament, and a the distance between the two cross linkers (Fig. 5D). λ is a Lagrange multiplier imposing that the length of the motor is equal to lm. In this subsection, we use for convenience the dynamics of the angle between motor and filaments ψ1 and ψ2 to characterise the orientation of the filaments. The dynamics in the limit where ψ1 and ψ2 relax quasi statically is given by 10 µ µ ∂x1 ∂t ∂x2 ∂t ǫµa2 ∂ψ1 ∂t ǫµa2 ∂ψ2 ∂t = f0 + = f0 + λ a λ a (x1 + lm cos ψ1 + x2 cos(ψ1 + ψ2)) (x2 + lm cos ψ2 + x1 cos(ψ1 + ψ2)) λ a λ a = − = − (x1lm sin ψ1 + x1x2 sin(ψ1 + ψ2)) (x2lm sin ψ2 + x1x2 sin(ψ1 + ψ2)) (A9) (A10) (A11) (A12) where ǫ ≪ 1 is a factor that vanishes in the quasi-static limit where the filaments rotate adiabatically. Solving for the Lagrange multiplier λ by imposing the constraint that the length of the motor is equal to lm, we obtain: λ = −ǫf0a3 (x1 + x2)(1 + cos(ψ1 + ψ2)) + lm(cos ψ1 + cos ψ2) A = (x1lm sin ψ1 + x1x2 sin(ψ1 + ψ2))2 + (x2lm sin ψ2 + x1x2 sin(ψ1 + ψ2))2 B = (x1 + lm cos ψ1 + x2 cos(ψ1 + ψ2))2 + (x2 + lm cos ψ2 + x1 cos(ψ1 + ψ2))2 A + Bǫa2 (A13) where the coefficient A vanishes when the filaments are aligned with the motor. Several situations can again be distinguished: • When the filaments are neither parallel nor antiparallel, from Eq. (A13), λ ∼ ǫ in the limit ǫ → 0. Eqs. (A9) and (A10) then yield at the lowest order in ǫ µ µ dx1 dt dx2 dt = f0 + O(ǫ) = f0 + O(ǫ) (A14) (A15) such that no steady-state exists in that situation. • When the filaments are parallel and point in the same direction, (ψ1, ψ2) = (0, π) or (ψ1, ψ2) = (π, 0). In that case λ = 0 and no steady-state exists for the motor, which runs on the two filaments before detachment. • When the filaments are antiparallel, (ψ1, ψ2) = (0, 0) (expansile configuration) or (ψ1, ψ2) = (π, π) (contractile configuration). In the expansile configuration, x1 + x2 + lm = χa, with χ = ±1. In the contractile configuration, x1 + x2 − lm = χa, with the same rule applying for χ. In both cases, Eq. (A13) yields λ = −χf0 and any position of the motor is a steady-state. To test for the stability of these steady states, we consider the dynamics of ψ1 and ψ2 around ψ1 = ψ2 = ψ∗ with ψ∗ ≡ 0 or π, ψ1 = ψ∗ + δψ1 and ψ2 = ψ∗ + δψ2. To regularise the dynamics around the parallel filaments state, we consider a situation with finite ǫ and consider perturbations verifying δψ ≪ √ǫ, such that λ ≃ −χf0. From Eqs. (A11) and (A12), we have then ǫµa ǫµa dδψ1 dt dδψ2 dt = = χf0 a χf0 a [(x1lm cos ψ∗ + x1x2)δψ1 + x1x2δψ2] [x1x2δψ1 + (x2lm cos ψ∗ + x1x2)δψ2] , (A16) (A17) so that the state ψ1 = ψ2 = ψ∗ is stable if χ(x1lm cos ψ∗ + x2lm cos ψ∗ + 2x1x2) < 0 and (x1lm cos ψ∗ + x1x2)(x2lm cos ψ∗ + x1x2) − x2 2 > 0, whereas it is unstable otherwise. Taking the initial condition to be x1 = x0/2 + (χa− lm cos ψ∗)/2 and x2 = −x0/2 + (χa− lm cos ψ∗)/2 at the steady state, the stability condition is satisfied when 1x2 χ(x2 0 + l2 m − a2) > 0 and x1x2χ cos ψ∗ > 0 . (A18) The results are summarized in the flow diagram in Fig. 5F. Red segments indicate the stable steady states. In the green region, there are no possible values of ψ1 and ψ2 for given values of x1 and x2. where θ is a fixed angle, and the Lagrange multiplier λ is obtained from the constraint that x2 µ µ ∂x1 ∂t ∂x2 ∂t = f0 − = f0 − λ lm λ lm (x1 − x2 cos θ) (x2 − x1 cos θ), (A20) (A21) 1 + x2 2− 2x1x2 cos θ = l2 m: (A22) λ = f0lm(x1 + x2)(1 − cos θ) (x1 − x2 cos θ)2 + (x2 − x1 cos θ)2 x1 = x2 = ± lm 2 sin θ 2 11 (A19) (A23) 4. Two filaments rigidly attached to the external network This situation corresponds to case (v). The mechanical potential U reads as for case (ii) U = −f0(x1 + x2) + λ(cid:18)qx2 1 + x2 2 − 2x1x2 cos θ − lm(cid:19) , with x1 and x2 the distances to the crossing point of the two filaments, taken positive in the direction of the filament towards which motors move (Fig. 5G). θ is the angle between the two filaments, such that n1.n2 = cos θ, with ni the unit vector giving the orientation of the filament i. λ is a Lagrange multiplier ensuring that the length lm of the motor is fixed. The dynamic equation for the motion of the motor on the two filaments Eq. (A1) then reads Two solutions can be found for the motor position on the two filament, assuming that they are not parallel (θ 6= 0 and θ 6= π): As pointed out in Ref. [7], a linear stability analysis around the steady-state indicates that only the positive solutions x1 = x2 = lm/[2 sin(θ/2)] is stable. The corresponding force dipole at equilibrium is positive and is given by d = lmf0 sin( θ 2 ) . (A24) Therefore, the dipole formed on two rigidly fixed, non-parallel filaments is always contractile (Fig. 5H-I). 5. Averaging force dipole To obtain the associated average force dipole exerted by the motor in these different configurations, we proceed as follows: denoting as c(0) f,2 the two angles giving the initial orientations of the filaments, x(0) f,2 the initial position of the two motors on the filaments, and rk (k = 1,··· , Nc with Nc the number of crosslinkers attached to the external network) the position of the crosslinkers on the filaments, we compute the motor-induced force dipole d reached at steady-state as shown in Fig. 5; cases (ii), (iii) and (v) by 12 the initial distance between the two filaments, θ(0) f,1 and θ(0) f,1, x(0) d(c(0) 12 , θ(0) f,1, θ(0) f,2, x(0) 1 , x(0) 2 ,{rk}k) = lmf0 2 cos ψ1 + lmf0 2 cos ψ2 , (A25) where the angles ψ1 and ψ2 between the motor and two filaments at steady-state are functions of c(0) x(0) 2 and {rk}k. Equation (A25) is obtained from the definition of the force dipole 12 , θ(0) f,1, θ(0) f,2, x(0) 1 , d = − lm 2 n · (f1 − f2) (A26) with n the unit vector giving the motor orientation pointing towards filament 1, and fi with i = 1, 2 are the forces exerted by the motor on the two filaments. The dependencies 1/ cos ψi (i = 1, 2) in Eq. (A25) arise from the condition that the projections of the forces fi on the filament i have magnitude f0 (Fig. 6). The average force dipole is then obtained by 0 < d >=Z dc(0) 12 Z 2π where the integration R d(x(0) a steady state, and Γ =R d(x(0) 1 , x(0) 2 )1. dθ(0) f,2 2π Z 2π 0 dθ(0) f,2 2π Z Nc Yk=1 drk Lf !Z d(x(0) 1 , x(0) 2 ) Γ d(c(0) 12 , θ(0) f,1, θ(0) f,2, x(0) 1 , x(0) 2 ,{rk}k) , (A27) 1 , x(0) 2 ) runs for all possible initial motor configurations x(0) 1 , x(0) 2 that eventually reach 12 FIG. 6. Schematic of the force dipole exerted by a motor. Each motor exerts a force of magnitude f0 parallel to the filament. The total force exerted by the motor on the filament are equal and opposite at steady-state and denoted f1 and f2. The contribution of the force normal to the filament arises from geometrical constrains. Overall, this results in a tension f = f1 = f2 acting within the motor, which is used here to define the force dipole exerted by a motor. Appendix B: Relationship between the stress and the average force dipole In this section, we discuss the two contributions to the stress generated by the filament and motor network, σf ij . We point out that in the absence of large scale deformations, σf σm elastic response of the filament network is non-linear. ij and ij can contribute to the total stress, when the 1. Two-dimensional, linear elastic material We consider here a two-dimensional elastic material subjected to the force dipoles exerted by motor filaments. The motor force dipoles consist of two opposite forces f k = ±f knk, separated by a distance lm, and where the index k label the motors. The motors induce a deformation in the elastic material. The total stress in the system is then the sum of the resulting elastic stress, denoted σf ij , and the stress generated within the motors, denoted σm ij : σij = σf ij + σm ij (B1) We start by discussing the average stress created by an ensemble of motors in a 2D material. Following Ref. [25], we now show that when the system is homogeneous, force balance on a section S of the network enclosed in a contour C allows to relate the stress generated within motors to the concentration and orientation of motors. To evaluate the stress, we first note that each motor k corresponds to a line under tension f k with length lm, orientation nk, centered at position xk, and joining the two points at coordinates xk − lm/2nk and xk + lm/2nk (Fig. 7A). Therefore, the two-dimensional stress field within the motors can be written σm ij (x) =Xk f knk i nk j Z lm 2 − lm 2 duδ(x − (xk + nku)) (B2) where u is a coordinate going along the line under tension f k. The average stress within a region S of surface area S A B 13 Motors are turned on FIG. 7. A. Schematic of a 2D elastic material, subjected to forces arising from force dipoles exerted by motor filaments. Motors result in a stress σij acting at the boundary of the box. B. Schematic of a 1D chain of non linear springs subjected to force dipoles. The chain contains N springs and is fixed at both ends. Motors result in a force σ along the chain. is then given by hσm ij i = dxσm ij (x) f knk i nk lm 2 − lm 2 j Z dxXk f klmnk i nk j duδ(x − (xk + nku)) (B3) (B4) (B5) 1 1 = S ZS S ZS S Xk = chdninji = 1 (B6) where c is the concentration of motors, dk = lmf k is the dipole strength of motor k, and the averaging h·i is performed over space. In a linearly elastic material, the average stress within the network of filaments is given by σf ij = 2E(cid:18)uij − 1 2 ukkδij(cid:19) + Kukkδij (B7) with E and K a shear and bulk elastic moduli, and uij = 1 stress generated in the elastic material in a region S with contour C and surface area S is then given by 2 (∂iuj + ∂jui) the gradient of deformation. The average dxσf ij hσf iji = = = 1 2E S ZS S ZS S ZC E dxuij + dxukkδij ZS S K − E S ZC E dlνiuj + dlνjui + K − E S ZC dlνkukδij (B8) (B9) (B10) 14 where ν is the vector normal to the contour C and dl an infinitesimal line element on the contour. If we consider a square box whose boundaries are fixed (Fig. 7A), or periodic boundary conditions, the contour integrals in Eq. (B10) vanish. Therefore, for a linear elastic material with fixed boundaries, the average stress arises entirely from the forces acting within the motors. This however does not apply to a non-linearly elastic material, as we show in the next section. 2. One-dimensional chain of non-linear springs We consider a simpler example in 1D of a periodic chain of N elastic springs. Each spring is located between positions xi and xi+1, with initial resting position xi = il. The two points at the end of the chain, x0 and xN , are not allowed to move. In addition, Nm motors are acting in parallel to a fraction n = Nm/N of the springs (Fig. 7B). The motor exerts a constant force f . The springs have a non-linear force-extension relation fe = k ∆l l l (cid:19)2 + k2(cid:18) ∆l (B11) with fe the force exerted by the spring, ∆l = xi+1 − xi − l the extension of the spring, and k and k2 are two spring constants. We assume that the springs are weakly non-linear, k2/k ≪ 1. In the initial resting position, ∆l = 0. The motors are then turned-on, driving a deformation of the the springs in the chain. The contraction of the springs which are in parallel with the motors is denoted −∆lm. Because the overall length of the chain is kept fixed, the deformation of the free springs is then given by ∆lmn/(1− n). We denote by σ the total force acting within the chain. Force balance imposes that the total force within the springs and the motors is fixed and equal to σ, giving σ = −k ∆lm l l (cid:19)2 + k2(cid:18) ∆lm + f = k ∆lm n l 1 − n + k2(cid:18) ∆lm l n 1 − n(cid:19)2 (B12) where the second part of the equality is the total force within the springs in parallel with a motor, and the third part the force within the free springs. Solving these equations and expanding to first order in f k2/k2 ≪ 1, one obtains or using the concentration of motors c = n/l, σ = f n + k2f 2n(1 − n) k2 σ = f lc + k2f 2 k2 cl(1 − cl) (B13) (B14) (B15) where σm = f lc is the average tension exerted within the motors. To leading order in the spring non linearity, k2 → 0, the force within the chain reduces to the average motor tension σm, in agreement with Eq. (B6). The non-linear elastic behaviour however brings a correction to this term σf = k2f 2cl(1 − cl)/k2, proportional to the motor force squared, f 2. = σm + σf Appendix C: Network relaxation time Crosslinker Filament FIG. 8. Schematic of a filament in the network. Crosslinkers bind to the filament with rate kcf on and unbind with rate 1/τc. We derive here an approximate expression for the network characteristic relaxation time. We consider a filament within the network, crossing other filaments that are themselves immobilised. We expect this last assumption to be valid for a large enough number of crosslinkers. The number of crossing points is assumed to largely exceed the total number of crosslinkers in the network. Crosslinkers bind and unbind the filament at crossing points with other filaments. A filament with two attached crosslinkers is completely fixed, and can only rotate when it has one attached crosslinker. For simplicity we consider here that the filament can not rearrange to relax stresses unless no crosslinker attaches it to other filaments in the network. Once a filament is free, it can move until a crosslinker binds to it and immobilise it. We therefore estimate the network relaxation time as the mean first passage time to a state where no crosslinkers bind the filament, from a state where one crosslinker binds the filament. 15 We consider the probability of having nc crosslinkers on the filament, P (nc). New crosslinkers bind to the filament on (Fig. 8), (Fig. 8). c with rate kcf onτc)]Nc/Nf , as crosslinkers bind to all filaments in the network with rate kc and each crosslinker binds two filaments. In addition, crosslinkers unbind from the filament with rate τ −1 The probability P (nc) then follows the master equation: on/(1 + kc on = 2[kc dP (nc) dt = nc + 1 τc P (nc + 1) + kcf onP (nc − 1) −(cid:18) nc τc + kcf on(cid:19) P (nc) d. Stationary distribution. The stationary distribution of Eq. (C1) is a Poisson distribution: P (nc) = e−kcf onτc (kcf onτc)nc nc! (C1) (C2) with mean and variance n∗ c = kcf onτc ≃ 2Nc/Nf for kc onτc ≫ 1. e. Mean first passage time. We now want to obtain the mean first passage time to reach a state where the filament has no bound crosslinker, nc = 0, starting from a configuration with n crosslinkers attached on the filament. We denote Tn this first passage time, and we follow a standard procedure to obtain its value [28, 29]. Tn satisfies the following equation for n > 0: Tn = kcf on kcf on + n/τc Tn+1 + n/τc kcf on + n/τc Tn−1 + 1 kcf on + n/τc (C3) where the first term corresponds to the probability of moving to a state with n + 1 bound crosslinkers, times the waiting time from the state n + 1, the second term is the product of the probability to move to a state with n − 1 bound crosslinkers times the waiting time from the state n − 1, and the last term is the average time spent in the state n. In addition, T0 = 0 by definition of the first passage time Tn, and we take a reflecting boundary condition at infinity, implying Tn − Tn−1 → 0 for n → ∞. Equation (C3) can be rewritten (Tn − Tn−1) = 1 kcf on(Tn − Tn+1) + (C4) n τc or defining zn = Tn+1 − Tn, zn = n kcf onτc zn−1 − 1 kcf on To solve this equation, one introduces Zn = zn(kcf onτc)n/n!, which satisfies then: Zn = Zn−1 − (kcf onτc)n kcf onn! Solving Eq. (C6) then yields the following expression for Zn, zn and Tn, using that Z∞ = 0: (kcf onτc)k kcf onk! Zn = ∞ Xk=n+1 zn = n! onτc)n (kcf ∞ Xk=n+1 (kcf onτc)k kcf onk! Tn = n−1 Xk=0 k! onτc)k (kcf ∞ Xm=k+1 (kcf onτc)m kcf onm! (C5) (C6) (C7) (C8) (C9) From this last expression, one finally obtains T1, ∞ (ekcf onτc − 1) c − 1) (en∗ (kcf onτc)m kcf onm! T1 = = = Xm=1 1 kcf on τc n∗ c 16 (C10) (C11) (C12) The time T1 therefore contains a factor increasing exponentially with the number of crosslinkers per filament n∗ c . We use the time T1 as an estimate for the network relaxation time. In Fig. 3f, we verify that the time T1 provides a good approximation for the relaxation time of the network. Appendix D: Effect of motor concentration and dispersion in motor forces ✁✂ . 2 0 . 1 0 . 0 0 . 5 0 . 4 0 . 3 0 . 2 0 . 1 0 . 0 0 ✄✂ 6 0 0 0 . 4 0 0 0 . 2 0 0 0 . 0 0 0 0 . 00 1 2 3 4 5 00 1 2 3 4 5 0 2 4 6 8 10 FIG. 9. A. Isotropic stress as a function of the number of motors Nm, for two values of myosin turn-over. In these graphs, we performed the simulations for the cases with identical motors (red circles) and with the a dispersion of 20% motor force and friction (blue triangles). Non-specified parameters are as in Fig. 3 of the main text. B. Isotropic stress as a function of the number of motors, in the absence of a cross linker. Green diamond marks, identical motors; no stress is generated. Blue triangle marks, a dispersion of 20% motor force and friction results in a non-zero stress for a high enough number of motors. Non-specified parameters are as in Fig. 3 of the main text. Stress was averaged here over 10τ after initialization. We have performed additional simulations measuring the isotropic stress σ as the concentration of myosin motors Nm/W 2 is varied (Figure. 9A). The reference stress σ0 is defined as being proportional to motor concentration, such that σ ∼ σ0 indicates that the stress is proportional to the myosin concentration. We find that σ has a weak non-linear dependence as a function of the myosin number Nm for the simulation parameters plotted in Figure 9A, and is nearly linearly increasing with Nm for slow enough turn-over. This indicates that the stress generated by the simulated network is roughly proportional to the myosin concentration. In our simulations, motors are identical and have the same stall force; as a result no stress is generated in the absence of cross linkers (Fig. 3 and Ref [30]). Introducing a dispersion in motor friction and stall force however can result in stress generation as the number of myosin motors is increased (Figure. 9B). The overall stress generated is much smaller than stresses generated for the same parameters with a cross linked network. Appendix E: Details of the model used for the numerical simulation 1. Main Components We introduce three components in our simulations: actin filaments, crosslinkers, and myosin filaments. Myosin filaments are represented by rigid rods with a length lm. To represent the attachment between myosin and actin filaments, two springs are added at the end of each myosin mini filament with stiffness kms and reference length 0, such that the tension in the end spring fms is fms = kmslms, with lms the length of the connecting spring. In practice kms is taken to be large enough that the spring maximum extension is very small compared to other lengths 17 in the simulation. The position of a myosin mini filament k on a filament i, relative to the center of mass of the filament i, is denoted ski. Actin filaments are treated as rigid rods with a finite length, lf . The position of the centre of mass of the filament is denoted xf,i and the unit vector giving the orientation of filament i is denoted nf,i. Crosslinkers are represented by springs connecting two actin filaments, with a stiffness kx and reference length 0, such that the tension with an crosslinker is fx = kxlx with lx the length of the crosslinker. In practice kx is chosen large enough such that the length of the cross linker is very small compared to other lengths in the simulation. We simulate the network in two dimensions x, y, in a periodic box of size W . Simulations are initialised by positioning filaments in the box with random positions and orientations. 2. Dynamics In the simulations, bond myosin minifilaments move on actin filaments. Myosin filaments detach from a filament when they reach the end of a filament. In addition, when myosin turnover is taken into account, they can spontaneously bind and unbind filaments. Crosslinkers can also bind and unbind filaments when crosslinker turnover is taken into account, and filaments can be removed and added in the network when filament turnover is taken into account. At every step of myosin motion, the network is relaxed quasi-statically to equilibrium. a. Myosin motion At every step, the end position of the myosin minifilament k on filament i, denoted ski, is updated according to the following equation: µ dski dt = f0 − fks · ni. (E1) where fks is the tension of the myosin end-spring connected to the filament, f0 is the active force generated by the myosin on the filament, and µ is an effective friction coefficient between the motor and the filament. The equation above is discretised with an Euler explicit scheme with time step dt. b. Filament and crosslinker turnover Crosslinkers are added with a rate kc of f = 1/τc. The number of bound cross linkers Ncb and unbound cross linkers Ncu add up to the total number of cross linkers Nc = Ncb + Ncu. At every step, each unbound cross linker has a probability of being added to the network kc ondt, and each bound cross linkers has a probability to be removed from the network kc of f dt. Unbound cross linkers are added by looking randomly for a free cross linking point on two filaments, and attaching the crosslinker there. on and removed with a rate kc Similar rules apply to turnover of myosin and actin filaments. The binding positions of two ends of the myosin filament are determined in the following way: Firstly, an actin filament and the position on it are randomly chosen, and one end of the myosin filament attached there. Then, the position of the second end of myosin filament is randomly chosen from all the attacheable positions, i.e. all the positions on actin filaments located lm away from the position on which the other end is attaching. c. Quasistatic relaxation At every time step, the network is relaxed quasi-statically to equilibrium. Below, we denote t∗ a fictitious time coordinate used for quasi-static relaxation. Quasi-static relaxation is performed by updating the filament center of mass ri and orientation ni = (cos θi, sin θi) according to the following equation: dri αt αr fik dt∗ =Xk dt∗ = ez · Xk dθi skni × fik! (E2) (E3) where fik is the force acting on filament i from the crosslinker or motor k. αt and αr are two translational and rotational fictitious friction coefficients, used for the quasi-static relaxation, and ez is the direction orthogonal to the plane of simulation. The same equations are used to iterate the position of myosin motors, rk and orientation nk. Iteration is performed by discretising equations (E2) and (E3) with an Euler explicit scheme, until the system reaches quasi-static equilibrium. In practice, a criterion must be used to specify when the system is close enough to equilibrium. To do so, the squares of the filament translational and rotational velocities are averaged according to: 18 e = 1 N "Xi (cid:18) dri dt∗(cid:19)2 + l2 f dt∗(cid:19)2 12Xi (cid:18) dθi dt∗(cid:19)2 +Xk (cid:18) drk + l2 m 12 Xk (cid:18) dθk dt∗(cid:19)2# (E4) where N is the total number of actin and myosin filaments. The quasi-static relaxation is completed when e is less than a threshold parameter emax. Appendix F: Supplementary movie legends • Supp. Movie M1 Simulation of a network with no crosslinker turnover, no filament turnover, Nc = 800 and τm = 100τ . Total simulation time, 1600τ . The number of cross linkers Nc is below the threshold for the network to exert a contractile stress. • Supp. Movie M2 Simulation of a network with no crosslinker turnover, no filament turnover, Nc = 1100 and τm = 100τ . Total simulation time, 1600τ . The number of cross linkers Nc is above the threshold for the network to exert a contractile stress. • Supp. Movie M3 Simulation of a network with no filament turnover, crosslinker turnover with τc = 100τ , Nc = 1200 and τm = 100τ . Total simulation time, 1600τ . The network collapses and does not exert a contractile stress in steady state. • Supp. Movie M4 Simulation of a network with filament turnover with τa = 100τ , crosslinker turnover with τc = 100τ , Nc = 800 and τm = 100τ . Total simulation time, 1600τ . The network reaches a steady-state where no contractile stress is exerted. • Supp. Movie M5 Simulation of a network with filament turnover with τa = 100τ , crosslinker turnover with τc = 100τ , Nc = 1200 and τm = 100τ . Total simulation time, 1600τ . The network reaches a steady-state where a contractile stress is exerted. [1] Guillaume Salbreux, Guillaume Charras, and Ewa Paluch. Actin cortex mechanics and cellular morphogenesis. Trends in cell biology, 22(10):536 -- 545, 2012. [2] Jean-Fran¸cois Joanny and Jacques Prost. Active gels as a description of the actin-myosin cytoskeleton. HFSP journal, 3(2):94 -- 104, 2009. [3] Mirjam Mayer, Martin Depken, Justin S Bois, Frank Julicher, and Stephan W Grill. Anisotropies in cortical tension reveal the physical basis of polarizing cortical flows. Nature, 467(7315):617 -- 621, 2010. [4] Martin Behrndt, Guillaume Salbreux, Pedro Campinho, Robert Hauschild, Felix Oswald, Julia Roensch, Stephan W Grill, and Carl-Philipp Heisenberg. Forces driving epithelial spreading in zebrafish gastrulation. Science, 338(6104):257 -- 260, 2012. [5] Tanniemola B Liverpool, M Cristina Marchetti, J-F Joanny, and J Prost. Mechanical response of active gels. EPL (Europhysics Letters), 85(1):18007, 2009. [6] Marina Soares e Silva, Martin Depken, Bjorn Stuhrmann, Marijn Korsten, Fred C MacKintosh, and Gijsje H Koenderink. Active multistage coarsening of actin networks driven by myosin motors. Proceedings of the National Academy of Sciences, 108(23):9408 -- 9413, 2011. [7] Nilushi L Dasanayake, Paul J Michalski, and Anders E Carlsson. General mechanism of actomyosin contractility. Physical review letters, 107(11):118101, 2011. [8] Martin Lenz, Margaret L Gardel, and Aaron R Dinner. Requirements for contractility in disordered cytoskeletal bundles. New journal of physics, 14(3):033037, 2012. [9] Martin Lenz, Todd Thoresen, Margaret L Gardel, and Aaron R Dinner. Contractile units in disordered actomyosin bundles arise from f-actin buckling. Physical review letters, 108(23):238107, 2012. [10] Simone Kohler and Andreas R Bausch. Contraction mechanisms in composite active actin networks. PloS one, 7(7):e39869, 2012. 19 [11] Jos´e Alvarado, Michael Sheinman, Abhinav Sharma, Fred C MacKintosh, and Gijsje H Koenderink. Molecular motors robustly drive active gels to a critically connected state. Nature Physics, 9(9):591 -- 597, 2013. [12] Minakshi Guha, Mian Zhou, and Yu-Li Wang. Cortical actin turnover during cytokinesis requires myosin ii. Curr Biol, 15(8):732 -- 6, Apr 2005. [13] Jean-Yves Tinevez, Ulrike Schulze, Guillaume Salbreux, Julia Roensch, Jean-Fran¸cois Joanny, and Ewa Paluch. Role of cortical tension in bleb growth. Proc Natl Acad Sci U S A, 106(44):18581 -- 6, Nov 2009. [14] K Tawada and K Sekimoto. A physical model of atp-induced actin-myosin movement in vitro. Biophys J, 59(2):343 -- 56, Feb 1991. [15] Y Imafuku, Y Emoto, and K Tawada. A protein friction model of the actin sliding movement generated by myosin in mixtures of mgatp and mggtp in vitro. J Theor Biol, 199(4):359 -- 70, Aug 1999. [16] Julicher and Prost. Cooperative molecular motors. Phys Rev Lett, 75(13):2618 -- 2621, Sep 1995. [17] Nobuhiro Morone, Takahiro Fujiwara, Kotono Murase, Rinshi S Kasai, Hiroshi Ike, Shigeki Yuasa, Jiro Usukura, and Akihiro Kusumi. Three-dimensional reconstruction of the membrane skeleton at the plasma membrane interface by electron tomography. J Cell Biol, 174(6):851 -- 62, Sep 2006. [18] Guillaume T Charras, Chi-Kuo Hu, Margaret Coughlin, and Timothy J Mitchison. Reassembly of contractile actin cortex in cell blebs. J Cell Biol, 175(3):477 -- 90, Nov 2006. [19] A Ott, M Magnasco, A Simon, and A Libchaber. Measurement of the persistence length of polymerized actin using fluorescence microscopy. Physical Review E, 48(3):R1642, 1993. [20] Ryoki Ishikawa, Takeshi Sakamoto, Toshio Ando, Sugie Higashi-Fujime, and Kazuhiro Kohama. Polarized actin bundles formed by human fascin-1: their sliding and disassembly on myosin ii and myosin v in vitro. Journal of neurochemistry, 87(3):676 -- 685, 2003. [21] Elizabeth W Kubalek, TQ Uyeda, and James A Spudich. A dictyostelium myosin ii lacking a proximal 58-kda portion of the tail is functional in vitro and in vivo. Molecular biology of the cell, 3(12):1455 -- 1462, 1992. [22] Svetlana Mukhina, Yu-Li Wang, and Maki Murata-Hori. Alpha-actinin is required for tightly regulated remodeling of the actin cortical network during cytokinesis. Dev Cell, 13(4):554 -- 65, Oct 2007. [23] Elizabeth M Reichl, Yixin Ren, Mary K Morphew, Michael Delannoy, Janet C Effler, Kristine D Girard, Srikanth Divi, Interactions between myosin and actin crosslinkers control Pablo A Iglesias, Scot C Kuo, and Douglas N Robinson. cytokinesis contractility dynamics and mechanics. Curr Biol, 18(7):471 -- 80, Apr 2008. [24] Martin Lenz. Geometrical origins of contractility in disordered actomyosin networks. Physical Review X, 4(4):041002, 2014. [25] R Aditi Simha and Sriram Ramaswamy. Hydrodynamic fluctuations and instabilities in ordered suspensions of self-propelled particles. Physical review letters, 89(5):058101 1 -- 058101 4, 2002. [26] Poul M Bendix, Gijsje H Koenderink, Damien Cuvelier, Zvonimir Dogic, Bernard N Koeleman, William M Brieher, Christine M Field, L Mahadevan, and David A Weitz. A quantitative analysis of contractility in active cytoskeletal protein networks. Biophysical journal, 94(8):3126 -- 3136, 2008. [27] Simone Kohler, Volker Schaller, and Andreas R Bausch. Structure formation in active networks. Nature materials, 10(6):462 -- 468, 2011. [28] CW Gardiner. Stochastic methods. Springer-Verlag, Berlin -- Heidelberg -- New York -- Tokyo, 1985. [29] Pedro A Pury and Manuel O C´aceres. Mean first-passage and residence times of random walks on asymmetric disordered chains. Journal of Physics A: Mathematical and General, 36(11):2695, 2003. [30] Martin Lenz, Margaret L Gardel, and Aaron R Dinner. Requirements for contractility in disordered cytoskeletal bundles. New journal of physics, 14(3):033037, 2012.
1504.07897
1
1504
2015-04-29T15:36:16
Conditions for positioning of nucleosomes on DNA
[ "physics.bio-ph", "q-bio.SC" ]
Positioning of nucleosomes along eukaryotic genomes plays an important role in their organization and regulation. There are many different factors affecting the location of nucleosomes. Some can be viewed as preferential binding of a single nucleosome to different locations along the DNA and some as interactions between neighboring nucleosomes. In this study we analyzed how well nucleosomes are positioned along the DNA as a function of strength of the preferential binding, correlation length of the binding energy landscape, interactions between neighboring nucleosomes and others relevant system properties. We analyze different scenarios: designed energy landscapes and generically disordered ones and derive conditions for good positioning. Using analytic and numerical approaches we find that, even if the binding preferences are very weak, synergistic interplay between the interactions and the binding preferences is essential for a good positioning of nucleosomes, especially on correlated energy landscapes. Analyzing empirical energy landscape, we discuss relevance of our theoretical results to positioning of nucleosomes on DNA \emph{in vivo.}
physics.bio-ph
physics
Conditions for positioning of nucleosomes on DNA Michael Sheinman, Ho-Ryun Chung Max Planck Institute for Molecular Genetics, 14195 Berlin, Germany (Dated: November 5, 2018) 5 1 0 2 r p A 9 2 ] h p - o i b . s c i s y h p [ 1 v 7 9 8 7 0 . 4 0 5 1 : v i X r a Positioning of nucleosomes along eukaryotic genomes plays an important role in their organization and regulation. There are many different factors affecting the location of nucleosomes. Some can be viewed as preferential binding of a single nucleosome to different locations along the DNA and some as interactions between neighboring nucleosomes. In this study we analyzed how well nucleosomes are positioned along the DNA as a function of strength of the preferential binding, correlation length of the binding energy landscape, interactions between neighboring nucleosomes and others relevant system properties. We analyze different scenarios: designed energy landscapes and generically dis- ordered ones and derive conditions for good positioning. Using analytic and numerical approaches we find that, even if the binding preferences are very weak, synergistic interplay between the in- teractions and the binding preferences is essential for a good positioning of nucleosomes, especially on correlated energy landscapes. Analyzing empirical energy landscape, we discuss relevance of our theoretical results to positioning of nucleosomes on DNA in vivo. PACS numbers: I. INTRODUCTION Our genome is packed and organized by nucleosomes -- histone octamers wrapped around by 147 bp of DNA [1, 2]. Nucleosomes are in some cases very well positioned while in others they are rather "smeared" along the DNA molecule [3, 4]. Their positioning properties are known to be important in the regulation of gene expression [5 -- 7]. There are many factors which determine positioning of nucleosomes along the DNA and their relative influence is a matter of active debate in the field (see Ref. [8] for a review). The most discussed positioning mechanism is the DNA sequence heterogeneity. It is well known that to wrap DNA around a nucleosome one needs different energies for different DNA sequences [9, 10]. In this case the de- bate is only about the importance of DNA sequence pref- erences, relative to other factors. An obvious competitor of sequences preferences for nu- cleosomes positioning is thermal fluctuations. All mea- sured binding energy differences between different se- quences do not exceed a few kBT even for specially de- signed strongest binders, which do not exist in known genomes [11 -- 15]. This indicates that, at least in equilib- rium, entropic forces are expected to play an important role. It is not entirely clear whether the nucleosomes reach (quasi)equilibrium and how they do it. It was sug- gested that some active chromatin remodeling enzymes, facilitate the equilibration of the nucleosomes by increas- ing the off-rate of the nucleosomes from the DNA [16]. These and others active chromatin remodeling enzymes [17, 18] and DNA-binding proteins [19] also affect nucle- osomes positioning by actively moving the nucleosomes and by DNA binding competition. In addition to external positioning signals, there are ar- guments for and evidences of interactions between neigh- boring nucleosomes along the DNA [20 -- 29]. The interac- tions are also expected to affect nucleosome positioning. This positioning factor is different from the one men- tioned above, since it depends not on an absolute position of a nucleosome on the DNA, but on a relative position of two nucleosomes -- the distance between two neighboring nucleosomes. In this study for simplicity we divide the positioning factors to two types. The first type includes all the fac- tors which determine the position of a single nucleosome. One can characterize it by an effective binding energy landscape of a nucleosome along the DNA molecule that depends only on the location of the nucleosome along the DNA. The second type corresponds to interactions between neighboring nucleosomes. We characterize this positioning factor by an effective interaction potential that depends only on the distance between two neigh- boring nucleosomes. We assume that nucleosomes can- not invade each others DNA territories (although it is not entirely true [30 -- 32] this is not expected to affect significantly the conclusions of this study). In addition we analyze only the equilibrium distribution of nucleo- somes, ignoring non-equilibrium aspects. In general, we address the following question: within the framework of the above assumptions, what should be the properties of the effective energy landscape and ef- fective interaction potential between neighboring nucle- osomes to achieve good positioning? We analyze energy landscapes with different properties and different interac- tion potentials and derive conditions leading to good po- sitioning of nucleosomes. We show that the interactions between nucleosomes can significantly improve their po- sitioning even on almost flat and highly correlated energy landscapes. In this case, if the positioning is good, one expects to observe also large length-scale fluctuations of nucleosome occupancy along the DNA. Comparing our results to empirical study, we find good qualitative and quantitative agreement. Before we start with detailed description of the model and its analysis, it might be instructive to illustrate the main message of the paper with a toy example. Con- sider non-correlated Gaussian binding energy landscape with standard deviation of 1.5kBT , on a circular DNA of length 400bp (see Fig. 1(a)). Ten non-interacting "nucle- osomes" of size 1bp cannot be well positioned with such a weak energetic disorder (see Fig. 1(c)). However, adding very strong interaction between the nucleosomes, such that the distance between them is restricted to 15bp, one get good positioning on the same, weak energetic profile (see Fig. 1(e)). Autocorrelation of an energy (see Fig. 1(b)) makes the positioning even more problematic (see Fig. 1(b)). However, again, strong interactions between the nucleosomes improves it to a reasonable level (see Fig. 1(d)). In the paper we will derive, within quite general set of assumptions, conditions for positioning on uncor- related and correlated correlated binding energy land- scapes. In Figs. 1(e) and (f) one can see that, when interactions between the nucleosomes are exploited for a better positioning, there are long length-scale fluctu- ations of the nucleosomes occupancy along the DNA -- there are long enriched and long depleted regions. Below we analyze the properties of such regions and demon- strate relevance of this effect to empirical data. The structure of the paper is as following. In Section II we formulate the model we use. In Section III we define the quantities we use to characterize positioning of nucleosomes on the DNA. In Section IV we analyze positioning of a single nucleosome. The purpose of this Section is not only didactic, because we use its results below. In Section V we analyze positioning of a many nucleosomes with only hard-core interactions. The pur- pose of this Section is to contrast it to the case with in- teractions between neighboring nucleosomes and demon- strate the importance of these interaction in Section VI. We generalize our results for energy landscapes with au- tocorrelation in Section VII. In Section VIII we discuss relevance of our conclusions to real systems and compare to empirical results. After discussion about tunability and robustness of positioning issues, emerging from our results, in Section IX, we summarize in Section X. We proceed now with a detailed description of our model. II. THE MODEL We analyze the following lattice based model, which is often used to calculate occupancy of DNA-binding pro- teins [33]. In a grand-canonical ensemble on average N nucleosomes are located on a linear DNA of length L, in units of bp with reflecting boundaries. Note, that near the saturation of the DNA by nucleosomes the grand-canonical and canonical ensembles may be differ- ent [18, 34]. Each nucleosome occupies W = 147bp on the DNA, such that if its leftmost position is bound to a site i another nucleosome cannot bind with its leftmost position to any of the sites in the interval [i − W + 1, i + W − 1]. Due to DNA sequence prefer- ences of nucleosomes or any other reason a nucleosome 2 bound with its leftmost position to site i possesses an energy Ei. In addition to this energy there is an interac- tion energy between two neighboring nucleosomes. Given the distance r ≥ 0 between the leftmost positions of the two nucleosomes the interaction energy is given by V (r). The hardcore interaction is realized by V (r) = ∞ for 0 ≤ r < W . To obtain the equilibrium properties of the nucleosomes we numerically solve the recursive equation for the partition function [35, 36]. Our focus is the following question: what should be the properties of the signal in the one-nucleosome energy profile along the DNA, Ei and the interaction between the nucleosomes, V (r), to achieve good positioning of nucleosomes on the DNA? In the next Section we define this question in more quantitative terms. III. QUANTITIES OF INTEREST In this paper we focus on several quantities which re- flect positioning of nucleosomes. Each one of them can be derived from an equilibrium probability of the site i to be covered by the leftmost position of a nucleosome, ni (start site probabilities). The average number of nu- cleosomes, N, is given by N = ni. (1) We also define an ordered vector of occupancies, no m, 1 is the occupancy of the most occupied site 2 is the occupancy such that no (site with the highest value of ni), no of the second-most occupied site etc. For cases when it is not important how a base-pair along the DNA is covered by a nucleosome (by which part of the nucleosome it is covered) the occupancy function L(cid:88) i=1 W−1(cid:88) j=0 L(cid:88) i=0 ρi = ni−j (2) is of an interest. We define the average coverage of the DNA by ρ = 1 L ρi = N W L . (3) There are different ways to define a measure of how well nucleosomes are positioned along the DNA. In this paper we use a very simple one: given that there are N nucleosomes, we define as P the fraction of nucleo- somes which occupies N most occupied locations along the DNA. Namely, N(cid:80) L(cid:80) m=1 P = m=1 N(cid:88) m=1 no m no m = 1 N no m. (4) 3 Figure 1: Illustration of how interactions between nucleosomes improves their positioning along the DNA. In this toy example "nucleosomes" are merely particles of 1bp size with disordered binding energy landscape on 400bp-long, circular "DNA". (a) Binding energies of a single nucleosome are i.i.d random variables with a Gaussian distribution with the standard deviation of σ = 1.5kBT . (b) Binding energies of a single nucleosome are normally distributed with the standard deviation of σ and are correlated, such that (cid:104)EiEi+r(cid:105) = σ2e rc , with rc = 20. (c) The probability that the site i is occupied by a nucleosomes, ni is plotted vs. i for 10 non-interacting nucleosomes located on the binding energy profile from (a). (d) The same as (c) but for the binding energy profile from (b). (e) The same as (c) but for nucleosomes with strong interactions, such that the distance between two neighboring nucleosomes is constrained to 15bp. (f) The same as (e) but for the binding energy profile from (b). − r In the case of a single nucleosome, N = 1 this definition becomes simply the occupation probability of the ground state -- the order parameter of, say, the Random Energy model [37]. For multiple nucleosomes P can be viewed as a fraction of well positioned nucleosomes. In the case of a perfect positioning P = 1, while in case of positioning whatsoever, P = N/L (cid:28) 1. The last quantity, N/L, can be at most 1/W for ρ = 1. As we show below, for correlated energy landscapes, in some cases nucleosomes are not positioned well on a sin- gle bp length scale but are positioned well within a few basepairs. In this case the value of P does not charac- terize fully the positioning goodness. For those case we exploit the following generalization of P. Denoting the profile around the m'th largest values of ni, as no m(s) we define for odd values of k N(cid:88) (k−1)/2(cid:88) no m(k). (5) Pk = 1 N m=1 s=−(k−1)/2 The value of Pk is the measure of positioning given that one does not care about fuzziness on the lengthscale of k. One can easily see that on the level of one bp resolution P1 is given by P. However, as we show below, on corre- non-correlated binding energy profile binding energy profile with correlation length of 20 10 "nucleosomes" with fixed distance, 15, between neighbors Probability that the site is occupied by a "nucleosome" 10 non-interacting "nucleosomes" (a) (b) (c) (d) (e) (f) lated energy landscapes and/or with interaction potential with wide wells the function Pk can be much more infor- mative than its single-bp resolution value P1 = P. We turn now to the consideration of positioning for different scenarios, starting from the simplest one. IV. POSITIONING OF ONE NUCLEOSOME It is instructive to consider first the simplest case of a single nucleosome, N = 1, on a DNA of a certain length, L. To position it on the DNA in equilibrium one should generate non-uniform energy profile along the DNA. Below we discuss possible energy profiles which can be roughly divided to designed and generic ones. A. Designed energy landscape Conceptually, the easiest way to position a nucleosome on a site j is to design an energy landscape such that up to an additive constant Ei=j = −E and Ei(cid:54)=j = 0. In this designed DNA case the positioning measure (4) is given by (we measure all energies in units of kBT ) P = no 1 = 1 1 + Le−E , (6) such that the nucleosome occupies site i with probability of order one if E (cid:38) ln L. Therefore, having in the arsenal only energies of the order of a few kBT one can position a single nucleosome only on short sequences of tens -- hundreds base-pairs, even on the best possible energy landscape. B. Disordered energy landscape with Gaussian distribution In the more generic case of a disordered energy land- scape the problem can be mapped to the Random Energy model [37 -- 40]. In this case the probability that the left- most location of the nucleosome is located on site i along the DNA is given by e−Ei Z where the partition function is ni = , L(cid:88) Z = e−Ei (7) (8) i=1 Therefore, to position a nucleosome on site i one has to fulfill the condition e−Ei ∼ Z. Consider the case where the energies {Ei} are a set of i.i.d random variables with a normal probability distribution with standard deviation σ: 4 Eo 1 −∞ In this case the lowest energy, Eo by 1, is well approximated The solution is given by √ −1 2erf 1 (cid:39) σ Eo Pr (E) dE = 1 L (10) (cid:19) − 1 (cid:18) 2 L √ (cid:39) −σ 2 ln L. (11) In the limit of zero temperature (or, equivalently infi- nite disorder strength σ) the nucleosome will occupy the state with the lowest energy (ground state) with proba- bility 1. However, for small non-zero temperatures non- ground states will be partly occupied, such that the occu- pation probability of the ground state is smaller than one. Consider the m-lowest energy (1-lowest energy means the lowest one, 2-lowest energy is the second lowest energy, etc.) on the DNA, Eo m. Its value can be well approxi- mated by Eo m −∞ Pr (E) dE = m L (12) − 1 (cid:19) (cid:18) 2m (cid:18) 2m L −1 (cid:19) − 1 L The solution is given by m (cid:39) σ Eo √ −1 2erf (cid:39) Eo 1 + σ σf ln m. (13) where the freezing disorder strength is given by √ σf (cid:39) − 2erf √ (cid:39) 2 ln L. (14) The affinities of the m-lowest state are given by m = e−Eo Ko m (cid:39) e−√ 2erf−1( 2m L −1) (cid:39) Ko mσ/σf 1 . (15) = (cid:32) (cid:33)−1 In the low temperature (high disorder) limit, σ > σf the occupation of the lowest state is given by (cid:16) σ P = no (cid:17) (cid:39) 21− σ − 1 1(cid:80) 1 (cid:39) Ko m Ko where ζ(s) =(cid:80)∞ (16) m=1 m−s is the Riemann zeta function. Above the freezing point σ < σf the sum in the equation above diverges. In this case the annealed approximation of the free energy is valid and the partition function is not widely distributed around its mean value 1 + 1 σf m ζ σ σf σf , Z (cid:39) (cid:104)Z(cid:105) = Leσ2/2. (17) In this regime the probability of occupation of any site is given by Pr (Ei) = i − E2 2σ2√ e 2πσ2 . (9) P = no 1 = m e−Eo Z (cid:39) eσ √ 2 ln L Leσ2/2 , (18) such that it vanishes in the thermodynamic limit, L → ∞. In genomes of lengths in the range L = 106 − 109 bp the freezing transition happens in the range σf (cid:39) √ 2 ln L = 5.3 − 6.4kBT . In sum positioning of a single nucleosome is determined by the disorder strength. It is well positioned on DNA of length L (such that P (cid:39) 1) with energetic Gaussian, uncorrelated disorder with width σ for σ (cid:29) √ 2 ln L and is "smeared" along the DNA (such that P (cid:28) 1) in the opposite limit of weak disorder, σ (cid:28) √ 2 ln L. In Fig. 2 In Appendix the above considerations are illustrated. A 1 we discuss positioning on energy landscape with non- Gaussian distributions. Now we turn to discuss position- ing of many nucleosomes on the DNA. A. Designed energy landscape 5 Consider first a designed case when there are N nucle- osomes of size W on a DNA of length L and N energy wells of energy −E while the rest of the DNA positions have zero energy. For the best positioning all the dis- tances have to be larger or equal to W , such that nu- cleosomes don't have to overlap to occupy all the energy wells. In this case to position well the nucleosomes one N . Then, having, say, 10 − 70 − 90% cov- needs E (cid:38) ln L erage of the DNA by nucleosomes of length W = 147, to position the nucleosomes one needs energy well to be deeper than 7− 4− 3kBT . Moreover, even if the wells are that deep but the number of nucleosomes differs from the number of energy wells the positioning is getting worse. To make the positioning more robust one can make the distance between the wells being random. However, do- ing this one has to keep the minimal distance between two neighboring wells to be W . Otherwise, the nucleosomes, being not able to overlap, spread more on the DNA de- creasing the positioning parameter, P (see Fig. 3). In a more generic, disordered case the positioning is more problematic. We turn now to discuss this case. Figure 2: Positioning of a single nucleosome on the DNA with length L = 106bp, such that the freezing transition is at σf (cid:39) 5.3kBT . Occupancy of the deepest energy well is plotted vs. disorder width. The dots represent numerical simulation -- median of 100 realizations of the disordered energy landscape energy. The lines represent the analytic solution: Eq. (16) for σ > σf and Eq. (18) for σ < σf. Inset: the same plot in the linear scale. V. POSITIONING OF MULTIPLE NUCLEOSOMES WITH ONLY HARDCORE INTERACTIONS Here we analyze positioning of N (cid:29) 1 nucleosomes with only hard-core interactions. The study of parti- cles with hard-core repulsions has a long history and is relevant to many applications. In the context of protein- DNA binding such a repulsion between proteins leads to crowding and influences the binding properties [41 -- 44]. Here we consider positioning of nucleosomes of a finite size, W ≥ 1 [45, 46] and focus on positioning properties on different energy landscapes. Figure 3: Numerical results for positioning of nucleosomes with only hard-core interactions with W = 147bp on designed energy landscapes of a length of L = 103 × W . Positioning parameter is plotted as a function of the coverage fraction ρ = N W/L. The energy profile is designed such that neigh- boring energy wells are separated by a distances with the following properties: Each distance is W + 63bp (circles), any distance is a sum of W and a number drawn from a geometric distribution with an average of 63bp (squares) and each dis- tance is a number drawn from a geometric distribution with an average of W + 63bp (diamonds). The depths of the energy wells relative to the rest positions on the DNA are: (a) 2, (b) 5 and (c) 8 kBT . The lines are to guide the eye. <05101520P10-610-510-410-310-210-11000510152000.20.40.60.81;00.20.40.60.81P00.10.20.30.40.50.60.70.80.91(c)8kBT(b)5kBT(a)2kBT B. Disordered energy landscape with Gaussian distribution VI. POSITIONING OF STRONGLY INTERACTING NUCLEOSOMES 6 (cid:113) Consider positioning of N nucleosomes on uncorre- lated disordered energy profile normally distributed with standard deviation σ (see Eq. (9)). In the regime σ (cid:28) N the nucleosomes are poorly positioned, while in 2 ln L the opposite regime, (cid:114) σ (cid:29) 2 ln L N , (19) the positioning is good. In sum, having, say, 10−70−90% coverage of the DNA by nucleosomes of length W = 147, to position the nucleosomes one needs disorder strength, σ to be larger than 3.8 − 2.9 − 2.4kBT . The derived requirement for positioning may sound weak. However, in fact it means that, say, for ρ = 70% and σ = 5kBT (moderate positioning regime, P (cid:39) 0.6, as shown in Fig. 4) the typical energy well for a nucle- osome is 14 ± 2kBT deep (see Eq. (11) with L replaced by L/N), relative to a random DNA sequence. In sum, one can see that without interactions the energy varia- tions required for a good positioning seem to be above the ones measured in experiments [11 -- 15]. In Appendix A 2 we discuss positioning of nucleosomes with only hard- core interaction on energy landscape with non-Gaussian distributions and show that the results in this Section do not change qualitatively in this case. In the next section we show how interactions between nucleosomes allow to position them with much weaker energetic disorder along the DNA. Consider N nucleosomes on a DNA of length L. As we show above, weak sequence specificity cannot posi- tion nucleosomes. In this Section we analyze positioning of interaction nucleosomes on designed and disordered energy landscapes. Interaction between neighboring nucleosomes, were suggested before as one of the driving forces, ordering nucleosomes [27 -- 29, 47]. Here, we consider for simplic- ity the minimal model of an interaction with an energy well when two neighboring nucleosomes are at a distance R − ∆ ≤ r ≤ R + ∆ from each other. Namely, the inter- action potential is of the form 0 e−V (r) = r < W ev = κ R + ∆ ≥ r ≥ R − ∆ ≥ W 1 else. (20) For simplicity we analyze narrow interaction energy well of only one bp, ∆ = 0. Further we discuss other possible potential functions in general and, in particular, impor- tance of a finite width of the interaction potential well, ∆ > 0. Figure 4: Positioning of nucleosomes with only hard-core in- teractions. Entropy of the DNA with length L = 103 × W and W = 147 as a function of the coverage fraction ρ = N W/L. The markers represent numerical simulation for σ = 2 (squares), 5 (circles), 8 (diamonds) and 11 (dots) kBT . Figure 5: neighboring nucleosomes used in the paper. Eq. (20) is presented. W = 147 and ∆ = 2. Illustration of the interaction strength between V (r) from In this particular case R = 154, For strong enough interactions strength (large values of κ) the nucleosomes gather to clusters, such that in each cluster the distance between the neighboring nu- cleosomes is R. As we show below this clustering effect can significantly improve positioning of nucleosomes. We demonstrate it first on a designed energy landscape. ;00.20.40.60.81P00.10.20.30.40.50.60.70.80.91rWRV(r)-90152"+1!0RW!11 A. Designed energy landscape B. Disordered energy landscape with Gaussian distribution 7 In order to exploit interactions between nucleosomes and position them on a weak but designed energy land- scape one should have a spatial resonance between the en- ergy wells distance and the preferable distance between neighboring nucleosomes, R. Consider first the case of periodic array of wells with energy −E, such that the affinity is K = eE, and set the number of nucleosomes to be the number of wells. Due to interactions nucleo- somes locally crystallize to ordered arrays with nearest- neighbors distance of R. Denoting the average length of a cluster by M, one gets the condition for the positioning of the cluster (and, therefore, all the nucleosomes in the cluster): or M E (cid:29) ln L N/M KM M (cid:29) L N . (21) (22) The average number of nucleosomes in a cluster, M, is given by M = 1 1 − Pr(r = R) , (23) where Pr(r = R) is the probability that the distance be- tween two neighboring nucleosomes is R. These quanti- ties can be calculated using the self-consistency equation: Pr(r = R) = 1 − 1 M = Kκ2 Kκ2 + L N/M . (24) √ In the case when the energy along the DNA is a set of independent normally distributed variables with stan- dard deviation σ (see Eq. (9)), local crystallization of interacting nucleosomes also plays a major role in posi- tioning for small values of σ. As shown in Section V B, with only hard-core interactions the positioning is possi- ble only when condition (19) holds. Here we discuss the opposite limit of weak disorder and show that interac- tions between the nucleosomes can position nucleosomes even in this case. Consider a cluster of crystallized nucle- osomes. The effective energy landscape for such a cluster possesses a stronger disorder than for an individual nucle- osome. Namely, for a cluster of, say, m nucleosomes the standard deviation of cluster's total energy distribution is mσ, where σ, as before, is the energy standard devi- ation of a single nucleosome. However, for m (cid:29) 1, this effective energetic disorder possesses an approximate pe- riodicity of R because shifting the cluster by this length the total energy of the cluster does not change much. Nevertheless, local crystallization, increasing the effective disorder strength relative to the one for a single nucle- osome, causes the positioning of clusters and, therefore, positioning of individual nucleosomes. Consider a single typical cluster of a size M (cid:29) 1. Typ- ical available space for it is given by L N/M , while the typ- √ 2 ln R. Therefore, ical minimal energy is given by it will be frozen if √ M σ √ √ M σ (cid:29) 2 ln R (26) and will be "smeared" on its available space in the oppo- site limit. The value of M can be estimated in the following way, using Eq. (23): Pr(r = R) = 1 − 1 M = κ κ + L N/M . (27) (cid:40)(cid:113) 1 κ N L M = κ (cid:29) L κ (cid:28) L N N . (28) Combining Eqs. (26) and (28), the required strength of interactions to position nucleosomes is given by W ρ (29) ln2 R. κ (cid:29) 4 σ4 In the case when σ (cid:29) √ 2 ln R condition (29) has to be replaced by κ (cid:29) L N . However, in this case condition (19) is satisfied, such that the positioning is possible with only hard-core interactions. Thus, as shown in Fig. 6, if at least one of the conditions (19) and (29) holds the positioning is good, such that P (cid:39) 1, while otherwise The solution is given by κ (cid:113) 1 K N L (cid:113) (cid:113) κ κ K N K N L (cid:29) 1 L (cid:28) 1 M = . (25) Thus, Therefore, satisfying condition (22), one gets strong improvement of the positioning (P (cid:39) 1) even with very weak wells K (cid:39) 1. However, if the wells do not have a good periodicity or their period is different from the preferential distance between the nucleosomes the inter- actions do not improve the positioning. In sum, one can position strongly interacting nucleo- somes on a designed energy landscape with shallow en- ergy wells. However, this positioning is not robust to change of nucleosome coverage fraction and requires fine tuning of the distances between energy wells along the DNA. We turn now to discuss positioning on generic en- ergy landscapes where the positioning is not so strong but is more robust to properties of the energy landscape and the interaction potential. P (cid:28) 1. One can see that strong interactions between nucleosomes are able to improve their positioning. In Appendix A 3 we discuss positioning of strongly interact- ing nucleosomes on energy landscape with non-Gaussian distributions and show that the results in this Section do not change qualitatively in this case. The described local clustering of nucleosomes not only improves positioning of nucleosomes but also has another consequences -- large length scale fluctuations of occu- pancy. We turn now to discuss this aspect of interactions- assisted positioning of nucleosomes. Figure 6: Positioning of nucleosomes with interactions on Gaussian energy landscape. Positioning parameter, P for the DNA with length L = 104 × W and W = 147 for the coverage fraction ρ = N W/L = 80% is plotted vs. disorder strength (left axis) and interaction strength (bottom axis) with prefer- able distance of R = 148. On the top one can see the av- erage size of the crystallized cluster of nucleosomes, derived from Eq. (28). On the right the typical binding energy of a nucleosome (relative to the average energy) is shown. The lines represent the analytic conditions for a good positioning, Eqs. (19) (dotted line) and (29) (solid line). C. Large-scale fluctuations of occupancy Apart length-scale, from the positioning of nucleosomes on small there is another feature that is highly influenced by interactions between neighboring nucleosomes -- the large-scale fluctuations of occupancy of nucleosomes. One can also interpret this effect as long length scale positioning. Without interactions oc- cupancy ρi averaged over thousands of base-pairs is not expected to deviate significantly from its average value, ρ. However, if neighboring nucleosomes possess a pref- erential distance, R which is smaller than the average linker length, W/ρ, and the nucleosomes are very well positioned one gets long DNA regions which are enriched by nucleosomes and, therefore, regions depleted with nu- cleosomes. The correlation length of the occupancy on enriched regions for strongly interacting and very well positioned nucleosomes is given (using Eq. (28)) by 8 (cid:114) (cid:18) W ρ RM ∼ R κ ρ W . (30) On this length scale the nucleosomes are enriched and their mean occupancy is given by W/R. The distance between the clusters of crystallized nu- cleosomes (with highly depleted occupancy) scales as (cid:19)(cid:114) L N/M − RM = − R κ ρ W . (31) In sum, strongly interacting and well positioned nucle- osomes are expected to exhibit highly fluctuating occu- pancy on large length-scales. In Section VIII we discuss how the described above considerations are relevant for more realistic energy landscapes and existing experimen- tal data, but before that we consider effects of another important feature of real systems -- autocorrelation of the energy landscape. VII. ENERGY LANDSCAPE WITH CORRELATIONS So far we discussed random, non-correlated energy landscapes. However, DNA sequence possesses correla- tions [48]. Moreover, even on a random DNA sequence an energy landscape is expected to be correlated for dis- tances smaller than 147 because small shifts of nucleo- somes along the DNA does not change completely the sequence covered by the nucleosome. This is why, as we discuss in the next Section, real energy landscape are ex- pected to possess certain autocorrelation. In this Section we discuss how the autocorrelation of the energy land- scape affects positioning of nucleosomes. We analyze the following scenario in this Section. The energy landscape is assumed to be Gaussian (see Eq. (9)) with an exponentially decaying autocorrelation, such that (cid:104)EiEi+r(cid:105) − r σ2 rc . = e (32) Here, rc (cid:29) 1 is the correlation distance, such that for distances much larger than rc the energies are nor corre- lated, while for distance much smaller than rc the varia- tion of energy is much smaller than σ. This model can be mapped to the generalized Random Energy model [49]. The condition for a good positioning on the single/few bp resolution, we derive below, correspond to the low- est/high temperature phase transition of that model, re- spectively. In this Section we assume that for rc = 0 the nucleo- somes are well positioned on the DNA. The correlation is an additional trouble for positioning and here we derive an additional condition for a good positioning in presence 9 imply that even for short-range autocorrelation of the energy profile with a realistic value of σ it impossible to position properly nucleosome on a single bp resolution without strong interactions between them. Figure 7: Positioning of nucleosomes with interactions on Gaussian energy landscape with an exponentially decaying autocorrelation with correlation coefficient rc = 20. Position- ing parameter, P for the DNA with length L = 104 × W and W = 147 for the coverage fraction ρ = N W/L = 80% is plot- ted vs. disorder strength (left axis) and interaction strength (bottom axis) with preferable distance of R = 148. On the top one can see the average size of the crystallized cluster of nucleosomes, derived from Eq. (28). On the right the typical binding energy of a nucleosome (relative to the average en- ergy) is shown. The line represents the analytic conditions for a good positioning, Eq. (35) or (36). However, the positioning on correlated energy land- scape is easier if one allow the nucleosome to be posi- tioned within a few bp. In order to see it we exploit the positioning function, defined in Eq. (5). This function, Pk, characterizes the positioning within the resolution of k base-pairs. In Fig. 9 one can see that in some cases even when the single bp resolution positioning is bad, P = P1 (cid:28) 1, the positioning within k = 3, 5, ... is signifi- cantly better. The condition for Pk to be of the order one is equivalent to condition (35) or (36) with rc replaced by rc/k. Namely, the condition for a good positioning within k bp is given by of the autocorrelation, on top of the conditions (19,29), for the non-correlated energy landscapes. We start from the simplest single-nucleosome case. Consider a single nucleosome on DNA of length L with an Gaussian energy landscape with standard devi- ation σ and exponential autocorrelation with correlation length rc (cid:29) 1. Conceptually we divide the DNA to L/rc "boxes" of length rc. In order to position the nucleo- somes in the box with the highest affinity one needs to . This condition is satisfied because we assume here that without autocorrelation the posi- 2 ln L. Thus, the problem is the positioning of the nucleosome within the box. satisfy σ (cid:29)(cid:113) tioning is good and, therefore, σ (cid:29) √ 2 ln L rc Within the box all the energies are highly correlated. A way to generate such a correlated energy landscape is to set [50] Ei+1 = e − 1 rc Ei + 1 − e − 2 rc Gi, (33) where Gi is an uncorrelated set of Gaussian random vari- ables with standard deviation σ. Thus, the standard de- viation of energies, Ei, if i is in the range much smaller than rc is given by . With such a standard devia- tion, to position a nucleosome in a box of size rc, one needs (cid:29) √ 2 ln rc or σ√ σ√ 2rc 2rc (cid:113) 4rc ln rc. (34) This is a strong constrain on the positioning. Even for rc = 5bp the positioning is bad unless σ is much larger than 6kBT . Condition (34) remains the same also for the case of non-interacting nucleosomes or nucleosomes with only hard-core interactions. This is because (34) does not de- pend on the length of DNA per nucleosome but only on the correlation distance of the energy landscape. This makes positioning of non-interaction (or interacting with only hard-core repulsion) extremely problematic. In the next Section we show that the value of rc is, at least, tens of base-pairs. For such a correlated energy landscape the positioning condition (34) is not expected to be satisfied. For interacting nucleosomes the standard deviation M σ√ of the effective energy landscape is given by , 2rc where M is the average number of nucleosomes in a crys- tallized cluster and given by Eq. (28). The positioning condition for strongly interacting nucleosomes is given by √ σ (cid:29)(cid:112) σ (cid:29) or (cid:114) 4rc M ln rc κ (cid:29) W ρ 16r2 σ4 ln2 rc. c or, equivalently, (35) (36) σ (cid:29) κ (cid:29) W ρ ln rc k (cid:114) 4rc 16(cid:0) rc M k k σ4 (cid:1)2 ln2 rc k . (37) (38) In Figs. 7 and 8 one can see the comparison of condition (35) or (36) to numerical results. The obtained results One can see in Fig. 8 that condition (37) (or (38)) can be much weaker than (35) (or (36)). In the next Section we study empirical landscape for which σ is roughly 1.5kBT and rc is roughly 100bp. Thus, if the interaction between nucleosomes is strong enough to crystallize them to clus- ters of size M = 10− 100 nucleosomes, one would expect to see bad positioning on the level of a single bp resolu- tion with P1 = 0.1− 0.25 ((35) does not hold) but within k = 9 bp the nucleosomes are positioned significantly better ((37) does hold) with P9 = 0.5 − 0.8 (see Fig. 9). In the next Section we discuss positioning properties on empirical energy landscape and, in general, relevance of the above considerations to real systems. Figure 8: Positioning of nucleosomes with interactions on Gaussian energy landscape with an exponentially decaying autocorrelation with correlation coefficient rc = 100. Posi- tioning parameter, P for the DNA with length L = 104 × W and W = 147 for the coverage fraction ρ = N W/L = 80% is plotted vs. disorder strength (left axis) and interaction strength (bottom axis) with preferable distance of R = 148. On the top one can see the average size of the crystallized cluster of nucleosomes, derived from Eq. (28). On the right the typical binding energy of a nucleosome (relative to the av- erage energy) is shown. The solid line represents the analytic conditions for a good positioning on a resolution of k = 1bp, Eq. (35) or (36). The dahsed line represents the analytic con- ditions for a good positioning on a resolution of k = 9bp, Eq. (37) or (38). VIII. RELEVANCE TO EMPIRICAL RESULTS So far we discussed positioning on artificial energy landscapes and dissected the phase diagram to different regimes. An obvious question to ask now is: where is the real system on the phase diagram? In this Section we try to get insight into this. To do so, we calculate an energy landscape using a model in Ref. [51]. In fact, there are many different models for a binding energy of a nucleo- some to a given sequence (examples include [51 -- 54], for review see, e.g., [55, 56]). We exploit only one of them, from Ref. [51], because we are not interested in predict- 10 ing locations of nucleosomes on some piece of DNA but in general properties of positioning of nucleosomes. In par- ticular, our goal in this section is to validate our results on artificial energy landscapes and show their relevance to more realistic energy landscapes. We start with contrasting the presented, artificial en- ergy landscapes and the one generated using the model in Ref. [51] on the genome of S. cerevisiae. First thing to note is that the distribution of the binding energies on the S. cerevisiae genome is close to, but deviates from a Gaussian (see Fig. 10). The standard deviation of the energy landscape is given by 1.6kBT . Interest- ingly, as shown in Fig. 10, the same model on a ran- domly shuffled S. cerevisiae genome yields significantly narrower energy landscape, well fitted by a Gaussian with σ = 1.24kBT . The energy landscape, generated by the model for Ref. [53] predicts even narrower energy land- scape (also shown in Fig. 10). Energy landscape with such a narrow distribution even with no autocorrelation is not expected to position well the nucleosomes without strong interactions (see Eq. (19) and Fig. 4). Interaction between nucleosomes can im- prove the positioning (see Eq. (29) and Fig. 6). However, as is expected, the calculated energy land- scape possesses certain autocorrelation. Some part of this autocorrelation is because shifting a nucleosome a few bp doesn't change entirely the bound sequence. This sort of autocorrelation, for distances smaller than 147bp exists even on a randomly shuffle genome (see Fig. 11). One can clearly see the periodic oscillations with an ap- proximate 10bp period [11, 57]. On top of that, due to some sequence correlation along the real, non-shuffled, S. cerevisiae genome, the autocor- relation of binding energy persist even for distances larger than 147bp (see Fig. 11). One reason for such a long-scale autocorrelation is that model in Ref. [51] is biased by GC content [58] and GC content possess significant autocor- relation along the S. cerevisiae genome [59] (see Fig. 12). The autocorrelation function clearly deviates from the simple exponential decay, which we assumed in our the- oretical considerations above. Roughly, the correlation distance is close to 100bp, making the positioning much more problematic, relative to uncorrelated energy land- scape with the same standard deviation (see Fig. 13(a) vs. (b)). On such an auto-correlated energy landscape, with only hardcore interactions, v = 0, nucleosomes are not well positioned on a single bp resolution (see a typical ni profile in Fig. 12). Namely, the single bp positioning parameter, P = 0.06 (calculated on the first chromo- some of S. cerevisiae), is much smaller than 1. With only hard-core interactions the peaks in ni are not only low, but also wide. In fact the positioning function Pk, shown in Fig. 13(a), demonstrates that the nucleosomes are "fuzzy" and poorly positioned even on the resolution of 10bp. The absence of good positioning can be partially at- tributed to strong autocorrelation of energy, because the 11 Figure 9: Positioning of nucleosomes with interactions on Gaussian energy landscape with an exponentially decaying autocor- relation with correlation coefficient rc = 100. Positioning parameter, Pk for the DNA with length L = 104 × W and W = 147 for the coverage fraction ρ = N W/L = 80% is plotted vs. k for different disorder strengths (left axis) and interaction strengths (bottom axis) with preferable distance of R = 148. The line represents the analytic conditions for a good positioning on a single bp resoltion, k = 1, Eq. (35) or (36). distribution of uncorrelated Gaussian energy landscape with σ = 1.6kBT results in much better positioning, as is discussed in Section V B. Indeed, calculation of posi- tioning of nucleosomes with only hard-core interaction on randomly shuffled energy landscape, calculated using model in Ref. [51] results in narrow (1bp) peaks of ni of an average height of P1 = P = 0.18 (see Fig. 13(b)). In contrast to the case with only hard-core interac- tions, the nucleosomes can be positioned much better in presence of interactions between neighboring nucleo- somes. For example, for the interaction strength of v = 9 and the preferable distance R = 154 the peaks of ni 12 the one used in Ref. [28] to fit qualitatively the 10n + 5 (or, sometimes, 10.6n + 8 [60]) periodicity found in many works, starting from Ref. [61]. In order to verify that our results do not depend qualitatively on the precise form of the interaction potential we used the same form as in Ref. [28], but with higher prefactor (12 instead of 5) and with a cutoff of 180bp for the computational purposes:  ∞ 12kBT cos 0 V (r) = (cid:16) 2π(r−W ) (cid:17) 10bp r < W 50bp W ≤ r ≤ 180 − r−W e . r > 180 Figure 10: Distribution of the binding energy of a nucleo- some on a S. cerevisiae (circles) and randomly shuffled S. cerevisiae genome (squares), calculated using the model in Ref. [51]. The standard deviation is equal to 1.6kBT for S. cerevisiae and 1.24kBT for randomly shuffled S. cerevisiae genome. The diamonds represent the distribution of the en- ergy landscape calculated in Ref. [53] for S. cerevisiae genome with a standard deviation of 0.8kBT . The lines are Gaussian fits with zero mean and respective standard deviations. (39) In Fig. 13(d) one can see that the interaction poten- tial in Eq. (39) is able to position nucleosomes on the energy landscape generated using the model in Ref. [51] within the resolution of 3 − 5bp. In the next Section IX we analyze in more detail robustness and tunability of positioning to different properties of the interacting potential. In sum, these results indicate that good position- ing of nucleosomes is possible even on a realistic en- ergy landscapes with narrowly distributed (small σ) and highly auto-correlated (large rc) energies, provided strong enough interactions between them. Beyond the positioning of nucleosomes on small length scales, as we described in Section VI C, strong interac- tions between neighboring nucleosomes change DNA oc- cupancy by nucleosomes on large length scales. We turn now to discuss how the theoretical predictions in Section VI C are relevant for realistic scenarios. A. Large-scale fluctuations of occupancy We start with nucleosomes with only hard-core inter- actions, as a reference case. In this case, since the energy profile is not well correlated for long distances, the oc- cupancy does not fluctuate significantly on large length scales (above a few nucleosome repeat lengths), as how in Figs. 12 and 14. However, as discussed in Section VI C, interactions locally crystallize nucleosomes, induc- ing large scale fluctuations in occupancy. In Figs. 12 (zoom on first 7· 104 bp of chromosome I of S. cerevisiae) and 14 (whole chromosome I of S. cerevisiae) one can see that interacting nucleosomes are distributed nonuni- formly along the DNA, in contrast to nucleosomes with only hard-core interactions. Interestingly, this sort of large length scale nonuniformity one also observes on the single-bp resolution data [62] (see Figs. 12 and 14). More- over, the calculated occupancy seems to follows quite con- sistently the experimental one on different length scales. This is especially surprising because the data from [62] is based on chemical cleavage, while the model used by use to calculate energy profile along the DNA was derived based on MNase digestion [51]. These results indicates that the interactions between the nucleosomes help to position them not only on the Figure 11: Autocorrelation of the binding energy of a nu- cleosome on a S. cerevisiae (black solid line) and randomly shuffled S. cerevisiae genome (dotted, red line), calculated us- ing the model in Ref. [51]. The inset is a zoom in on the main plot. are much higher, such that the positioning parameter is P1 = 0.3 (see Figs. 12 and 13(c)). The width of the peaks is 3− 5bp, such that the positioning function is Pk ∼ 0.45 for k ≥ 3, as can be seen in Fig. 13(c). Can one position nucleosome with a more realistic in- teraction potential? A reasonable choice seems to be E-10-50510Pr(E)10-610-410-2100r050100150200250300C(r)00.10.20.30.40.50.60.70.80.91r0102030C(r)0.40.60.81 13 Figure 12: Analysis of the first 7 · 104bp of S. cerevisiae's first chromosome. The lines on top of the figure depict exons. In panel number (I) the GC content is plotted. (II) Binding energy profile calculated using model from Ref. [51]. The additive constant term is such that the mean energy along the chromosome is zero. (II) Calculation of the nucleosome distribution for interacting nucleosomes with R = 154 and v = 9. The chemical potential is such that ρ = 0.8. The lines with dots represent ni values, while the thick, red line represent occupancy level. The thin black line is the occupancy from the data in Ref. [62]. Both the occupancies are smoothed, such that only 100 lowest Fourier modes are presented. (III) The same as for the (II) panel but with only hard-core interactions, v = 0. short length-scales (a few bp), but also on the long length-scales, inducing large long-scale fluctuations in nu- cleosomes occupancy. In other words, strong interactions between the nucleosomes naturally yields long nucleo- somes diluted and enriched regions along the genome. We summarize this Section with the following con- clusions. On a realistic energy landscape nucleosomes are much better localized in presence of strong interac- tions. In this case the calculated results, predicting large length scale occupancy fluctuations, agree qualitatively and, surprisingly, quantitatively with the experimental data. We turn now to a more detailed analysis of how properties of the interaction potential affect distribution of nucleosomes along the genome. IX. TUNABILITY AND ROBUSTNESS OF THE POSITIONING Several relevant question are still to be answered. How robust are the obtained results for the made assump- tions? What if the interaction potential possesses a cer- tain width (∆ > 1 in Eq. (20))? How robust the positions of nucleosomes to change of parameters, like the strength of the interactions v, preferable distance, R, number of wells, and the width of the interactions potential, ∆. Can one "tune" the distribution of nucleosomes along the DNA changing (locally or globally) the parameters of the interaction potential? We addressed some of these issues above, considering 14 (39) Figure 13: Positioning goodness in different cases. In each panel the left plots are the average profile of the N largest values of ni (insets are zoom out of the main plots). The right plots are the positioning function Pk vs. k. (a) Energy landscape from Ref. [51] with only hard-hore interactions between the nucleosomes, v = 0. (b) The same as in (a) but the energy landscape (not the sequence) is randomly shuffled. (c) The same as in (a) but nucleosomes interact with v = 9, R = 154, ∆ = 0. (d) The same as in (a) but nucleosomes interact with the potential from Eq. (39). (e) The same as in (c) but nucleosomes interact with ∆ = 2. (f) The same as in (a) but nucleosomes interact with two wells potential, R = 154 and R = 164. more realistic potential between the nucleosomes, in Eq. (39), and found that the conclusions are quite robust to a particular form of the potential. In this Section we ad- dress these questions more systematically. We focus here on the energy profile from the previous Section, calcu- lated using model from Ref. [51] and take as a starting point potential with a single well with R = 154, v = 9, ∆ = 0. In this case the positioning is reasonably good (see Figs. 12 and 13(c)). We start by changing the width of the interaction po- tential, ∆. Before we always (except from analyzing the potential form from Eq. (39)) assumed that the interac- tion potential is sharp to the level of a single bp, ∆ = 0. Of course this assumption is not realistic and real effec- tive interaction potentials probably possess a finite width and more than one energy well [20, 63, 64]. However, changing the width to a few base-pairs does not change qualitatively the results. Taking ∆ = 2 the positioning remains good on the length scale of 2∆+1 = 5, such that the typical peak of ni has a height of P (cid:39) 0.3 and width of 5bp, as shown in Fig. 13(e). As one can see in Fig. 15, the occupancy on large length scales does not change much as one tune the value of ∆. Even on the level of a single bp resolution the Pearson correlation coefficient of ρi for a potential with a width of one bp (∆ = 0) and 5bp (∆ = 2) is 0.87. The precise locations of nucleosomes do change, however. Looking on the N = ρL/W largest values of ni for both values of ∆ we observe an overlap of 18%. This value makes sense because we expect that, upon changing the width of the potential well from 1 bp to 5bp, 20% of the nucleosomes will remain in their posi- tions and 80% will move ±2 bp, within the new potential well. In sum, widening of the potential width does not change the distribution of nucleosomes on a large scale, but makes their position uncertain on the length scale of 2∆ + 1. We add now one more well to the interaction potential at 10bp from the first one, R = 154 + 10 = 164. The positioning gets slightly worse, such that the typical peak of ni has a height of P (cid:39) 0.2 and width of 3bp, as shown 15 Figure 14: Occupancy of nucleosomes on the global scale of the whole first chromosome of S. cerevisiae. Thin blue line represents calculation of the nucleosome distribution for interacting nucleosomes with R = 154 and v = 9. Thick, gray line is the occupancy from the data in Ref. [62]. The calculated occupancy for nucleosomes with only hard-core interactions is represented by the thin, red, almost flat line. The chemical potential set for calculations is such that ρ = 0.8. All the occupancies are smoothed, such that only 50 lowest Fourier modes are presented. Figure 15: Changing the width of the potential, ∆. Thick red line represents calculation of the nucleosome distribution for interacting nucleosomes with R = 154, v = 9 and ∆ = 0. Thin, blue line represents calculation of the nucleosome distribution for interacting nucleosomes with R = 154, v = 9 and ∆ = 2, such that the width of the potential well is 2∆ + 1 = 5bp. The chemical potential set for calculations is such that ρ = 0.8. All the occupancies are smoothed, such that only 100 lowest Fourier modes are presented. Locally the occupancy values are also highly correlated. The Pearson correlation coefficient of ρi (without any smoothing) for ∆ = 0 and 2 is 0.87. The values of ρi for ∆ = 0 and 2 differ by 19% on average. However, the locations of nucleosomes for ∆ = 0 and 2 significantly differ: N = ρL/W locations with the highest values of ni for the two cases possess an overlap of only 18%. i#10500.511.52;i00.40.81.20=0datai#10500.511.52;i00.51 in Fig. 13(f). As one can see in Fig. 16, the occupancy on large length scales does not change much as one tune the value of ∆. Even on the level of a single bp resolution the Pearson correlation coefficient of ρi for a potential with one and two wells is 0.75. The precise locations of nucleosomes do change, dramatically, however. Looking on the N = ρL/W largest values of ni for both values of ∆ we observe an overlap of only 3%. In sum, another well in the potential width does not change the distribution of nucleosomes on a large scale, but does change their positions on small length scale. The depth of the potential well does affect the goodness of positioning and has to be strong enough to have any ef- fect. However, once a reasonable positioning is achieved, its value does not change things much. As one can see in Fig. 17, the occupancy on a large length scales does not change much as one tune the value of v from 9 to 11. Even on the level of a single bp resolution the Pear- son correlation coefficient of ρi for a potential with one and two wells is 0.94. Even the precise locations of nu- cleosomes do not change, dramatically. Looking on the N = ρL/W largest values of ni for both values of v we observe an overlap of 67%. As one would expect, the position of the well, R, does not affect significantly the positioning properties and dis- tribution of the nucleosomes on a large scale. As one can see in Fig. 18, the occupancy on large length scales does not change much as one tune the value of R from 154 to 160. On the level of a single bp resolution the Pearson correlation coefficient of ρi for a potential with one and two wells is 0.62. However, precise positioning of nucle- osomes is very sensitive to the value of R: N = ρL/W locations with the highest values of ni for the two cases of R = 154 and R = 160 possess an overlap of only 1%. We conclude that, in principle, a cell, tuning the prefer- able distance between nucleosomes, can control their dis- tribution along some part of DNA without changing sig- nificantly large scale properties, like an average position- ing goodness and large length scale occupancy. X. SUMMARY In this article we focus on goodness of positioning of nucleosomes on the DNA. We make several simplifying assumptions. We assume that we can take into account all the positioning factors by have an effective energy landscape and an effective interaction potential between neighboring nucleosomes. We ignore that nucleosomes can invade each others DNA territories. In addition we analyze only the equilibrium distribution of nucleosomes, ignoring very probable non-equilibrium aspects of nucle- osomes positioning. However, even within this simplified framework we clarify a few aspects of nucleosome posi- tioning, which do not seem to depend on these details. Looking on a generic energy landscape with some en- ergy distribution width, some energy typical autocorre- lation distance and some interaction potential between 16 neighboring nucleosome we derive condition for a good positioning. We briefly summarize the conditions in the following paragraph. Assuming that neighboring nucleosomes possess a preferable distance with an affinity κ (cid:29) 1, relative to other distances, the number of locally crystallized nucle- osomes, M (cid:29) 1, is given by Eq. (28), where N is the average number of nucleosomes and L is the length of the DNA. The the positioning is expected to be good on an uncorrelated Gaussian disorder with standard de- viation σ is condition Eq. (26) holds. On a correlated energy landscape with a certain correlation distance, rc, the positioning conditioning depends on the required res- olution. For a good positioning within k bp it is given by Eq. (37). Importantly, without strong interactions, κ (cid:39) 1, the conditions do not seem to hold for realistic parameters, indicating an important role of effective interactions be- tween the nucleosomes in their positioning. If the posi- tioning, as we suggest, is controlled by interactions, one expect to see long length-scale fluctuations of nucleosome occupancy. This conclusion agrees with empirical data on occupancy of nucleosomes. Moreover, the derived large length-scale occupancy profile, derived from an effective energy landscape and interaction potential which is suffi- cient to position the nucleosomes is similar to the empiri- cal one. We also analyze the robustness of positioning to parameters of the model. The parameters can vary with time and be different on different parts of the genome. In fact, as we demonstrate, tuning some parameters one can dramatically change the distribution of nucleosomes. In sum, our study emphasizes an important role of inter- action between the nucleosomes and indicates range of parameters needed for it. We expect this knowledge to be important for better understanding of organization of our epigenome. Appendix A: Non-Gaussian energy landscapes The distribution of the binding energies along the DNA does not have to be Gaussian. Here we analyze two other possible scenarios for the energy landscapes. 1. Positioning of one nucleosome Here we discuss positioning of a single nucleosome on up-exponential and down-exponential energy landscapes. a. Disordered energy landscape with down-exponential distribution Consider scenario with what we denote as down- exponential distribution: Pr(Ei) = 1 E e EiE ; −∞ < Ei ≤ 0 ; E > 0. (A1) 17 Figure 16: Changing the number of wells of the potential. Thick red line represents calculation of the nucleosome distribution for interacting nucleosomes with R = 154, v = 9 and ∆ = 0. Thin, blue line represents calculation of the nucleosome distribution for interacting nucleosomes with two wells (both with ∆ = 0) R = 154 and R = 164, v = 9. The chemical potential set for calculations is such that ρ = 0.8. All the occupancies are smoothed, such that only 100 lowest Fourier modes are presented. Locally the occupancy values are also highly correlated. The Pearson correlation coefficient of ρi (without any smoothing) for one and two wells is 0.75. The values of ρi for one and two wells differ by 18% on average. However, the locations of nucleosomes for one and two wells significantly differ: N = ρL/W locations with the highest values of ni for the two cases possess an overlap of only 3%. Figure 17: Changing the strength of the potential, v. Thick red line represents calculation of the nucleosome distribution for interacting nucleosomes with R = 154, v = 9 and ∆ = 0. Thin, blue line represents calculation of the nucleosome distribution for interacting nucleosomes with R = 154, v = 11 and ∆ = 0. The chemical potential set for calculations is such that ρ = 0.8. All the occupancies are smoothed, such that only 100 lowest Fourier modes are presented. Locally the occupancy values are also highly correlated. The Pearson correlation coefficient of ρi (without any smoothing) for v = 9 and 11 is 0.94. The values of ρi for v = 9 and 11 differ by 15% on average. In this case even the locations of nucleosomes for v = 9 and 11 do not significantly differ: N = ρL/W locations with the highest values of ni for the two cases possess an overlap of 67%. In this case the typical minimal energy can be estimated using Eq. (10) to be 1 (cid:39) −E ln L. Eo (A2) The partition function be be estimated separately in two regimes (two phases in the thermodynamic limit): in the nonfrozen regime, E (cid:28) 1 the partition function is given by The positioning parameter is this case is given by P (cid:39) LE−1 (cid:28) 1, (A4) such that the positionig is poor for E (cid:28) 1. In the opposite regime E (cid:29) 1 the integral in Eq. (A3) diverges and the partition function is dominated by the deepest wells. Thus, it can be estimated using k-minimal energies, given by Z (cid:39) L 0 −∞ e−E Pr(E)dE = L 1 − E . (A3) k (cid:39) E ln Eo L k . (A5) i#10500.511.52;i00.51i#10500.511.52;i00.51 18 Figure 18: Changing the well location of the potential, R. Thick red line represents calculation of the nucleosome distribution for interacting nucleosomes with R = 154, v = 9 and ∆ = 0. Thin, blue line represents calculation of the nucleosome distribution for interacting nucleosomes with R = 160, v = 9 and ∆ = 0. The chemical potential set for calculations is such that ρ = 0.8. All the occupancies are smoothed, such that only 100 lowest Fourier modes are presented. Locally the occupancy values are also highly correlated. The Pearson correlation coefficient of ρi (without any smoothing) for R = 154 and 160 is 0.62. The values of ρi for R = 154 and 160 differ by 23% on average. On the resolution of a few bp the locations of nucleosomes for R = 154 and 160 are very different: N = ρL/W locations with the highest values of ni for the two cases possess an overlap of only 1%. Therefore, (cid:18) L (cid:19)E k ∞(cid:88) k=1 Z (cid:39) = LE ζ(E), (A6) such that the positioning parameter in this regime is given by P (cid:39) 1 ζ(E) . and is larger than 1/2 and close to one for E (cid:29) ζ−1(2) (cid:39) 1.7. (A7) (A8) In sum, for down-exponential distribution of energies the only requirement for a good positioning is that the parameter E is larger than one for any length of the DNA. b. Disordered energy landscape with up-exponential distribution Consider another scenario with what we denote as down-exponential distribution: 2. Positioning of multiple nucleosomes with only hardcore interactions Here we consider positioning of nucleosomes with hard-core interactions on up-exponential and down- exponential energy landscapes. a. Disordered energy landscape with down-exponential distribution Consider positioning of N nucleosomes on uncorrelated disordered energy profile down-exponentially distributed with some parameter E (see Eq. (A1)). In the regime E (cid:28) 1 the nucleosomes are poorly positioned, while in the opposite regime E (cid:29) 1, the positioning is good. The derived requirement for positioning may sound weak. However, in fact it means that, say, for ρ = 70% and E = 1.5kBT (moderate positioning regime, P (cid:39) 0.6) the typical energy well for a nucleosome is 7.5 ± 2kBT deep (see Eq. (A2) with L replaced by L/N), relative to a random DNA sequence. Pr(Ei) = (A9) b. Disordered energy landscape with up-exponential distribution 1 E e− EiE ; 0 ≤ Ei < ∞ ; E > 0. 2 (cid:39) Eo 1 (cid:39) Eo In this case there are many energies which are close to 3 (cid:39) ... (cid:39) 0. Positiong is impossible zero, Eo for such energy landscape; for any value of E one has P (cid:28) 1. However, as we show below interaction between nucleosomes can induce reasonable positioning even on such a energy landscape. Consider positioning of N nucleosomes on uncorrelated disordered energy profile up-exponentially distributed with some parameter E (see Eq. (A9)). In this case, as for a single nucleosomes, the positioning is poor for any value of E. i#10500.511.52;i00.51 3. Positioning of strongly interacting nucleosomes Here we discuss positioning of strongly interacting nu- cleosomes on up-exponential and down-exponential en- ergy landscapes. a. Disordered energy landscape with down-exponential distribution b. Disordered energy landscape with up-exponential distribution 19 Consider positioning of N nucleosomes on uncorrelated disordered energy profile down-exponentially distributed with some parameter E (see Eq. (A1)). In this case a cluster of m (cid:29) 1 crystallized nucleosomes has a Gaussian mE. Us- energy landscape with a standard deviation of ing the same arguments as for the Gaussian disorder one can derive two conditions for a good positioning. The first is for positioning of weakly interacting nucleosomes, such that M (cid:39) 1. In this case the condition is given by Eq. (A8). √ If this condition is not satisfied one needs strongly in- teracting nucleosomes, such that √ √ ME (cid:29) 2 ln R or, using Eq. (28), κ (cid:29) 4 E 4 L N ln2 R. (A10) (A11) Consider positioning of N nucleosomes on uncorrelated disordered energy profile up-exponentially distributed with some parameter E (see Eq. (A9)). In this case, as for a single nucleosomes, the positioning is poor for any value of E and one needs strongly interacting nucleosomes for a good positioning. If nucleosome strongly interact and form crystallized clusters of M (cid:29) 1 nucleosomes on average (and this hap- pens when κ (cid:29) L N ), the positioning condition is given by Eq. (A11). Thus the condition for positioning in this case is κ (cid:29) L N and 4 E 4 L N ln2 R. (A12) [1] R. D. Kornberg and Y. Lorch, Cell 98, 285 (1999). [2] T. J. Richmond and C. A. Davey, Nature 423, 145 (2003). [14] T. E. Takasuka and A. Stein, Nucleic acids research p. arXiv:0805.4017 (2008). gkq279 (2010). [3] D. J. Gaffney, G. McVicker, A. A. Pai, Y. N. Fondufe- Mittendorf, N. Lewellen, K. Michelini, J. Widom, Y. Gi- lad, and J. K. Pritchard, PLoS genetics 8, e1003036 (2012). [4] E. C. Small, L. Xi, J.-P. Wang, J. Widom, and J. D. Licht, Proceedings of the National Academy of Sciences 111, E2462 (2014). [5] C. Jiang and B. F. Pugh, Nature Reviews Genetics 10, [15] D. D. Winkler, U. M. Muthurajan, A. R. Hieb, and K. Luger, Journal of Biological Chemistry 286, 41883 (2011). [16] R. Padinhateeri and J. F. Marko, Proceedings of the Na- tional Academy of Sciences 108, 7799 (2011). [17] M. Vignali, A. H. Hassan, K. E. Neely, and J. L. Work- man, Molecular and cellular biology 20, 1899 (2000). [18] V. B. Teif and K. Rippe, Nucleic acids research 37, 5641 [6] L. Bai and A. V. Morozov, Trends in genetics 26, 476 [19] P. Korber, T. Luckenbach, D. Blaschke, and W. Hörz, 161 (2009). (2010). [7] Y. Belch, J. Yang, Y. Liu, S. A. Malkaram, R. Liu, J.- J. M. Riethoven, and I. Ladunga, PloS one 5, e12984 (2010). [8] K. Struhl and E. Segal, Nature structural & molecular biology 20, 267 (2013). [9] E. Trifonov, Nucleic acids research 8, 4041 (1980). [10] S. C. Satchwell, H. R. Drew, and A. A. Travers, Journal of molecular biology 191, 659 (1986). [11] A. Thåström, P. Lowary, H. Widlund, H. Cao, M. Ku- bista, and J. Widom, Journal of molecular biology 288, 213 (1999). [12] A. Thåström, P. Lowary, and J. Widom, Methods 33, 33 (2004). [13] A. V. Morozov, K. Fortney, D. A. Gaykalova, V. M. Studitsky, J. Widom, and E. D. Siggia, arXiv preprint (2009). ences 89, 1095 (1992). view E 70, 011915 (2004). view E 74, 031919 (2006). Molecular and cellular biology 24, 10965 (2004). [20] J. Widom, Proceedings of the National Academy of Sci- [21] B. Mergell, R. Everaers, and H. Schiessel, Physical Re- [22] F. Mühlbacher, H. Schiessel, and C. Holm, Physical Re- [23] R. Stehr, N. Kepper, K. Rippe, and G. Wedemann, Bio- physical journal 95, 3677 (2008). [24] J.-P. Wang, Y. Fondufe-Mittendorf, L. Xi, G.-F. Tsai, E. Segal, and J. Widom, PLoS computational biology 4, e1000175 (2008). [25] S. A. Grigoryev, G. Arya, S. Correll, C. L. Woodcock, and T. Schlick, Proceedings of the National Academy of Sciences 106, 13317 (2009). [26] Y. Liu, C. Lu, Y. Yang, Y. Fan, R. Yang, C.-F. Liu, 20 biology 414, 749 (2011). [27] R. V. Chereji, D. Tolkunov, G. Locke, and A. V. Moro- zov, Physical Review E 83, 050903 (2011). [28] R. V. Chereji and A. V. Morozov, Journal of statistical physics 144, 379 (2011). [29] D. A. Beshnova, A. G. Cherstvy, Y. Vainshtein, and V. B. Teif, PLoS computational biology 10, e1003698 (2014). [30] M. Engeholm, M. de Jager, A. Flaus, R. Brenk, J. van Noort, and T. Owen-Hughes, Nature structural & molec- ular biology 16, 151 (2009). [31] W. Möbius, B. Osberg, A. M. Tsankov, O. J. Rando, and U. Gerland, Proceedings of the National Academy of Sciences 110, 5719 (2013). [32] R. V. Chereji and A. V. Morozov, Proceedings of the National Academy of Sciences 111, 5236 (2014). [33] V. B. Teif and K. Rippe, Journal of Physics: Condensed Matter 22, 414105 (2010). [34] C. P. Woodbury, Biopolymers 20, 2225 (1981). [35] G. Gurskii and A. Zasedatelev, Biofizika 23, 932 (1977). [36] V. B. Teif and K. Rippe, Briefings in bioinformatics 13, 187 (2012). [37] B. Derrida, Physical Review Letters 45, 79 (1980). [38] U. Gerland, J. D. Moroz, and T. Hwa, Proceedings of the National Academy of Sciences 99, 12015 (2002). [39] M. Slutsky and L. A. Mirny, Biophysical journal 87, 4021 [40] M. Sheinman, O. Bénichou, Y. Kafri, and R. Voituriez, Reports on Progress in Physics 75, 026601 (2012). [41] J. D. McGhee and P. H. von Hippel, Journal of molecular biology 86, 469 (1974). [42] G.-W. Li, O. G. Berg, and J. Elf, Nature Physics 5, 294 (2004). (2009). (2012). [43] M. J. Morelli, R. J. Allen, and P. R. Ten Wolde, Bio- physical journal 101, 2882 (2011). [44] M. Sheinman and Y. Kafri, Physical biology 9, 056006 [45] I. R. Epstein, Biophysical chemistry 8, 327 (1978). [46] R. D. Kornberg and L. Stryer, Nucleic acids research 16, 6677 (1988). [47] S. Lubliner and E. Segal, Bioinformatics 25, i348 (2009). [48] C. Peng, S. Buldyrev, A. Goldberger, S. Havlin, F. Sciortino, M. Simons, H. Stanley, et al., Nature 356, 168 (1992). [49] B. Derrida, Journal de Physique Lettres 46, 401 (1985). [50] M. Deserno (2010). [51] N. Kaplan, I. K. Moore, Y. Fondufe-Mittendorf, A. J. Gossett, D. Tillo, Y. Field, E. M. LeProust, T. R. Hughes, J. D. Lieb, J. Widom, et al., Nature 458, 362 (2009). [52] M. Y. Tolstorukov, V. Choudhary, W. K. Olson, V. B. Zhurkin, and P. J. Park, Bioinformatics 24, 1456 (2008). [53] G. Locke, D. Tolkunov, Z. Moqtaderi, K. Struhl, and A. V. Morozov, Proceedings of the National Academy of Sciences 107, 20998 (2010). [54] T. van der Heijden, J. J. van Vugt, C. Logie, and J. van Noort, Proceedings of the National Academy of Sciences 109, E2514 (2012). [55] D. Tolkunov and A. V. Morozov, Advances in protein chemistry and structural biology 79, 1 (2010). [56] V. Teif, A. Shkrabkou, V. Egorova, and V. Krot, Molec- ular Biology 46, 1 (2012), ISSN 0026. [57] E. N. Trifonov, Physics of Life Reviews 8, 39 (2011). [58] H.-R. Chung, I. Dunkel, F. Heise, C. Linke, S. Krobitsch, A. E. Ehrenhofer-Murray, S. R. Sperling, and M. Vin- Figure 19: Positioning of nucleosomes with interactions on down-exponential energy landscape. Positioning parameter, P for the DNA with length L = 104×W and W = 147 for the coverage fraction ρ = nW/L smaller than 80% is plotted vs. disorder strength (left axis) and interaction strength (bottom axis) with preferable distance of R = 148. On the top one can see the average size of the crystallized cluster of nucleosomes, derived from Eq. (28). On the right the typical binding energy of a nucleosome (relative to the average energy) is shown. The lines represent the analytic conditions for a good positioning, Eqs. (A8) (dotted line) and (A11) (solid line). Figure 20: Positioning of nucleosomes with interactions on up-exponential energy landscape. Positioning parameter, P for the DNA with length L = 104 × W and W = 147 for the coverage fraction ρ = nW/L smaller than 80% is plotted vs. disorder strength (left axis) and interaction strength (bottom axis) with preferable distance of R = 148. On the top one can see the average size of the crystallized cluster of nucleosomes, derived from Eq. (28). On the right the typical energy peak (not well) is shown on the length of L/N. The line represents the analytic conditions for a good positioning, Eq. (A12). N. Korolev, and L. Nordenskiöld, Journal of molecular gron, PloS one 5, e15754 (2010). [59] K. R. Bradnam, C. Seoighe, P. M. Sharp, and K. H. Wolfe, Molecular biology and evolution 16, 666 (1999). [60] A. B. Cohanim, Y. Kashi, and E. N. Trifonov, Journal of Biomolecular Structure and Dynamics 23, 559 (2006). [62] K. Brogaard, L. Xi, J.-P. Wang, and J. Widom, Nature [63] F. Strauss and A. Prunell, The EMBO journal 2, 51 486, 496 (2012). (1983). [61] D. Lohr and K. Van Holde, Proceedings of the National zov, Physical Review E 83, 050903 (2011). Academy of Sciences 76, 6326 (1979). [64] R. V. Chereji, D. Tolkunov, G. Locke, and A. V. Moro- 21
1511.04270
1
1511
2015-11-13T13:00:50
Dynamic curvature regulation accounts for the symmetric and asymmetric beats of Chlamydomonas flagella
[ "physics.bio-ph", "q-bio.CB", "q-bio.SC" ]
Axonemal dyneins are the molecular motors responsible for the beating of cilia and flagella. These motors generate sliding forces between adjacent microtubule doublets within the axoneme, the motile cytoskeletal structure inside the flagellum. To create regular, oscillatory beating patterns, the activities of the axonemal dyneins must be coordinated both spatially and temporally. It is thought that coordination is mediated by stresses or strains that build up within the moving axoneme, but it is not known which components of stress or strain are involved, nor how they feed back on the dyneins. To answer this question, we used isolated, reactivate axonemes of the unicellular alga Chlamydomonas as a model system. We derived a theory for beat regulation in a two-dimensional model of the axoneme. We then tested the theory by measuring the beat waveforms of wild type axonemes, which have asymmetric beats, and mutant axonemes, in which the beat is nearly symmetric, using high-precision spatial and temporal imaging. We found that regulation by sliding forces fails to account for the measured beat, due to the short lengths of Chlamydomonas cilia. We found that regulation by normal forces (which tend to separate adjacent doublets) cannot satisfactorily account for the symmetric waveforms of the mbo2 mutants. This is due to the model's failure to produce reciprocal inhibition across the axes of the symmetrically beating axonemes. Finally, we show that regulation by curvature accords with the measurements. Unexpectedly, we found that the phase of the curvature feedback indicates that the dyneins are regulated by the dynamic (i.e. time-varying) component of axonemal curvature, but not by the static one. We conclude that a high-pass filtered curvature signal is a good candidate for the signal that feeds back to coordinate motor activity in the axoneme.
physics.bio-ph
physics
Dynamic curvature regulation accounts for the symmetric and asymmetric beats of Chlamydomonas flagella Pablo Sartori1,(cid:89), Veikko F. Geyer2,(cid:89), Andre Scholich1, Frank Julicher1, Jonathon Howard2,† 1 Max Planck Institute for the Physics of Complex Systems, Dresden, Germany 2 Department of Molecular Biophysics and Biochemistry, Yale University, New Haven, Connecticut (cid:89)These authors contributed equally to this work. †[email protected] Abstract Axonemal dyneins are the molecular motors responsible for the beating of cilia and flagella. These motors generate sliding forces between adjacent microtubule doublets within the axoneme, the motile cytoskeletal structure inside the flagellum. To create regular, oscillatory beating patterns, the activities of the axonemal dyneins must be coordinated both spatially and temporally. It is thought that coordination is mediated by stresses or strains that build up within the moving axoneme, but it is not known which components of stress or strain are involved, nor how they feed back on the dyneins. To answer this question, we used isolated, reactivate axonemes of the unicellular alga Chlamydomonas as a model system. We derived a theory for beat regulation in a two- dimensional model of the axoneme. We then tested the theory by measuring the beat waveforms of wild type axonemes, which have asymmetric beats, and mutant axonemes, in which the beat is nearly symmetric, using high-precision spatial and temporal imaging. We found that regulation by sliding forces fails to account for the measured beat, due to the short lengths of Chlamydomonas cilia. We found that regulation by normal forces (which tend to separate adjacent doublets) cannot satisfactorily account for the symmetric waveforms of the mbo2 mutants. This is due to the model's failure to produce reciprocal inhibition across the axes of the symmetrically beating axonemes. Finally, we show that regulation by curvature accords with the measurements. Unexpectedly, we found that the phase of the curvature feedback indicates that the dyneins are regulated by the dynamic (i.e. time-varying) component of axonemal curvature, but not by the static one. We conclude that a high-pass filtered curvature signal is a good candidate for the signal that feeds back to coordinate motor activity in the axoneme. Author Summary The swimming of many microorganisms is powered by periodic bending motion of cilia and flagella. The flagellar beat results from a feedback: dynein motors generate sliding forces that bend the flagellum; and bending leads to deformations and stresses, which feed back and regulate motors. Three alternative feedback mechanisms have been proposed: regulation by the sliding forces, regulation by the curvature of the axoneme, 1 and regulation by the normal forces that tend to separate adjacent doublets. In this work we combine theoretical and experimental techniques to test whether any of these mechanisms can account for the waveforms of the short flagella of the unicellular alga Chlamydomonas. We show that the sliding control mechanism can not produce bend propagation for short flagella, which results in a poor fit to the data. Comparison of the waveforms of wild type Chlamydomonas with those of a mutant that has a symmetric beat argues against normal force regulation. By contrast, the waveforms predicted by the curvature control model accord with the experimental data. Importantly, we make the surprising prediction that the motors respond to the time derivative of curvature, rather than curvature itself, hinting at an adaptive mechanism within the cilium. Introduction Cilia and flagella are long, thin organelles whose regular oscillatory bending waves propel cells through fluids and drive fluid flows across the surfaces of cells. The internal motile structure, the axoneme, contains nine doublet microtubules, a central pair of single microtubules, motor proteins in the axonemal dynein family and a large number of additional structural and regulatory proteins [1]. The axonemal dyneins power the beat by generating sliding forces between adjacent doublets. The sliding is then converted to bending by constraints at the base of the axoneme (e.g. provided by the basal body) and/or along the length of the axoneme (e.g. nexin links) [2 -- 4]. While the constrained-sliding mechanism of bend formation is well established, it is not known how the activities of the dyneins are coordinated in space and time to produce the periodic beating pattern. It is thought that the beat is the result of feedback. The axonemal dyneins generate forces that deform the axoneme; the deformations, in turn, regulate the dyneins. Because of the geometry of the axoneme, deformation leads to stresses and strains that have components in various directions (e.g. axial and radial). However, which component (or components) regulates the dyneins is not known. Three different, but not mutually exclusive, models for dynein regulation have been suggested in the literature, see Fig. 1. According to the sliding control model, dyneins are regulated by tangential forces acting parallel to the long axis of the microtubule doublets [5 -- 9]. According to the curvature control model, dyneins are regulated by doublet curvature [10 -- 12]. According to the normal force control model, dyneins are regulated by transverse forces that act to separate adjacent doublets when they are curved [13, 14]. Which of these mechanisms regulates the beat of the axoneme is not known. In this work, we isolated Chlamydomonas axonemes, and tracked them with high spatial and temporal resolution. Performing a Fourier decomposition of the data, we observed that the beat of wild type and mbo2 mutant axonemes have similar dynamics. The primary difference lies in the static asymmetry characteristic of the wild type beat, which is lacking in the mutant [15]. By developing a two-dimensional theory for the axonemal beat in the presence of a static asymmetry, we show that sliding control, curvature control, and normal force control can all, in principle, give raise to wave propagation in asymmetric axonemes. However, the observed beats of wild type cells and mbo2 mutants are not consistent with sliding and normal force control. By contrast, curvature control accorded with the wild type and mbo2 beats provided that the curvature signal adapts (or is insensitive to) the static component of curvature. 2 Figure 1. Opposing filaments model of the axoneme and dynein regulation. (A) Scheme of two opposing filaments bent by motors, as seen in the bending plane xy. The two filaments are constrained to have a spacing a0. The dyneins step towards the base of the doublets. Dyneins sitting at the bottom filament with their head (blue circle) on the top filament produce a tensile force density +f on the top filament, which tends to slide it towards the distal end; and a compressive force density −f on the bottom filament. The dyneins sitting on the opposite filament create antagonistic forces. The local sliding displacement is given by ∆, and the sliding at the base is ∆b. The spring and dashpot at the base correspond to the compliance of the base with stiffness kb and friction coefficient ξb. The green springs indicate the normal compliance which supports a normal force f⊥. The position of the point at the arc length s from the origin is r, characterized by a tangent vector t, a normal vector n, and a tangent angle ψ with respect to the horizontal axis of the lab-frame xy. (B) Schematic of dynein regulation mechanisms. Under curvature control the dynein head detaches due to an increase in curvature. In sliding control detachment is enhanced by a tangential loading force, and in normal force control it is the normal force that enhances detachment. Signs indicate doublets polarity. Theoretical model Periodic beat of an asymmetric cilium We model the axonemes imaged in our experiments as two opposing inextensible filaments immersed in a fluid (see Fig. 1). In this two-dimensional model, the filaments are connected by motor proteins, which are oriented in both directions and so can slide filaments in either direction. This captures the crucial idea that motors on opposite sides of the axoneme generate antagonistic bending forces [16]. Elastic elements keep the filaments at a constant distance a0 from each other. The position and shape of the filament pair is given at each time by the vector r(s), a function of the arc length s along the centerline between the filaments. Calculating the tangent vector as t(s) = r(s), where dots denote arc length derivatives, allows us to define the tangent angle ψ(s) with respect to the horizontal axis of the lab-frame that characterizes the shape of the filament at each time. For a given filament shape, the pair of filaments will have a local sliding displacement ∆(s) with respect to the other [17 -- 19]. The sliding is related to the shape of the axoneme via ∆(s) = ∆b + a0(ψ(s) − ψ(0)) , (1) where ∆b is the basal sliding and a0 is the spacing between the filaments (see Fig. 1). To calculate how the shape of the axoneme depends on the active motor forces f , as well as the passive mechanical properties of the axoneme (such as its bending 3 ABsliding controlnormal force controlcurvature controldetachdetachdetachloadforcenormalforce+++ rigidity κ) and the viscous properties of the fluid, we establish a force balance, see Appendix. To calculate the mechanical forces that the axoneme exerts on the fluid we introduce a work functional U that contains the effect of axonemal rigidity and active motors, and differentiate with respect to r. This force is balanced by the friction force of the fluid, proportional to the velocity of the axoneme at each point, and so we have − Π· ∂tr = δU/δr, where Π is the friction matrix [20]. For a slender body at low Reynolds number we have Π = ξnnn + ξttt, with ξn and ξt the normal and tangential hydrodynamic friction coefficients of the cilium and n(s) a unit vector normal to t(s) (see Fig. 1). Force balance provides non-linear equations of motion which are given in the Appendix. We then look for periodic solutions of these equations at the beat frequency of the axoneme. For observed periodic beats, we can decompose the tangent angle into Fourier modes ψ(s, t) = ψn(s) exp(inωt) , (2) ∞(cid:88) n=−∞ where ω is the fundamental angular frequency of the beat (ω/2π is thus the beat frequency), t is time, and ψn are the harmonics which satisfy ψ−n = ψ∗ n to keep the angle real. We refer to n = 0 as the static mode and to n = 1 as the fundamental mode. The same decomposition can be done for the sliding force f and the tension τ along the cilium. For each value of n, we obtain an equation of motion. The n = 0 equation allows us to calculate the time-averaged shape. The n = 1 equation allows us to calculate the motion at the beat frequency. where F0 is the zeroth mode of the integrated sliding force F = −(cid:82) L The time-averaged shape is calculated from the static force balance a0F0 = κ ψ0, s f (s(cid:48))ds(cid:48). As we will show later, Chlamydomonas axonemes have a static curvature C0 that is approximately ψ0(s) = C0 [15, 21]. According to the static force balance, constant along the arc length, bending the axoneme with a constant curvature requires accumulation of forces at the distal end [22]. The static component of the sliding force density is thus f0 = −δ(s − L)κC0/a0 , (3) where δ is the Dirac delta function, and the minus sign comes from the fact that dynein is a minus end directed motor which has a negative sliding velocity. See Appendix for details on the sign convention. To obtain the dynamic motion at the beat frequency, we expand the non-linear dynamic equations on the dynamic modes. Keeping the linear term we obtain the following equation of motion for the fundamental mode (n = 1) i¯ωψ1 = −.... 0 τ1 = −(1 + ¯ζ) ¯C0( ¯ζ τ1 − ¯C 2 ... ψ 1 − f1) ψ 1 + f1 + (1 + ¯ζ) ¯C0 τ1 + ¯ζ ¯C 2 0 ( ψ1 − f1) , (4) which has constant coefficients and has to be supplemented by appropriate boundary conditions. Here all quantities have been made dimensionless as indicated in Appendix. There are three dimensionless constants noted with overbars: the dimensionless static curvature ¯C0 = C0L; the normalized friction ¯ζ = ξn/ξt; and the normalized frequency ¯ω = ωξnL4/κ, which is sometimes referred to as the "sperm number" [20]. Equation 4 shows how an oscillatory active sliding force f1 produces dynamic bending of the cilium under appropriate parameter and boundary conditions. Equation 4 is the equation of motion for an asymmetrically beating axoneme. It generalizes the equation for a symmetric axoneme [6,8,12], for which the static curvature is C0 = 0 and the axial tension τ1 = 0. For C0 (cid:54)= 0 new terms appear, and the system of equations is of order six rather than four in the symmetric case. The magnitude of the new 4 terms can be estimated by considering the plane wave limit, in which ψ1 = exp(−2πis/λ), with λ the wavelength. For example, the fourth term in the right hand side of the equation is of order ¯ζ ¯C 2 0 (2π)2/λ2 and is in phase with the first term, that is also present in the symmetric theory and is of order (2π)4/λ4. For Chlamydomonas axonemes λ ∼ L and C0 ∼ π/L, and since ζ ≈ 2 [23, 24] the ratio of these terms is ∼ 0.5, thus the new contributions due to the asymmetry can not be neglected. A similar reasoning shows that the tension term is in anti-phase, and its contribution is of order ∼ 1. This shows that the additional terms are non-neglibible. Thus, for the large observed asymmetry of the Chlamydomonas axoneme, we expect that in general coupling between the n = 0 and n = 1 modes significantly modifies the dynamics of the beating axoneme. Surprisingly, we will later show that this is not the case for beats regulated by curvature. Three mechanisms of motor control Equation 4 shows how an oscillating sliding force can produce a dynamic beating pattern. However, we do not expect the motor proteins themselves to be oscillators that drive axonemal motion at the beat frequency. Rather, we expect that the sliding forces generated by the dyneins are regulated directly or indirectly by the shape of the axoneme through feedback. In this view, the motor force produces a bend, and the bend regulates the motor forces. Here we consider three different types of feedback motor regulation: sliding control, in which motors are regulated by sliding of the filaments ∆; curvature control, in which they are regulated by the curvature ψ; and normal force control, in which motors are regulated by the normal force f⊥ that keeps the filaments spacing at a0, see Fig. 1. The sliding, curvature, and normal force can be decomposed in Fourier modes just as the angle ψ. The fundamental mode of the sliding, ∆1, relates to that of the angle, ψ1, and the basal sliding, ∆b,1, through Eq. 1, which is linear. Thus in sliding and curvature control the feedback is linear. In the case of the normal force control, however, the feedback from shape is non-linear, see Appendix, and the fundamental mode is given by f⊥,1 = C0(F1 + κ ψ1/a) , (5) where F1 is the fundamental mode of the integrated sliding force. This expression vanishes for the symmetric case in which ψ0 = 0, indicating that in symmetric beats the normal force only has higher harmonics. Thus the asymmetry C0 (cid:54)= 0 of the Chlamydomonas cilium opens a way to regulation by normal forces, something impossible in symmetrically beating cilia of which bull sperm is an approximate example [8]. The most general expression for the dependence of the sliding force on sliding, curvature and normal force is fn = χ(nω)∆n + β(nω) ψn + γ(nω)f⊥,n , (6) where n > 0 is a mode-index, and χ(ω), β(ω) and γ(ω) are complex frequency dependent coefficients describing the response to sliding, curvature and normal forces, respectively. These are the generalization to active motors of the familiar linear response of passive systems. For example, the basal force of the axoneme is given by Fb = kb∆b + ξb∂t∆b, and thus the fundamental mode of the basal force is given by Fb,1 = χb(ω)∆b,1 with χb(ω) = kb + iωξb. Similarly, the coefficients χ, β and γ will depend on active motor properties. Equation 6 expresses how changes in the shape of the axoneme affect the motor force, while Eq. 4 shows how the motor force influences the axonemal shape. Together, they form a dynamical system which can become unstable and produce spontaneous oscillations [6, 25]. At the critical point these oscillations are perfectly periodic and we only retain the fundamental mode from Eq. 6. Finally, we note that for n = 0 the response coefficients must vanish, because the observed constant static 5 curvature implies that the static component of the force vanishes all along the length except at the very distal end of the axoneme (see Eq. 3). While for now we will take this for granted, we will later demonstrate that this is indeed the case for beats resulting from dynamic curvature control. Figure 2. High precision tracking of isolated axonemes Panel A corresponds to Chlamydomonas wild type axonenes, panel B to mbo2 mutant axonemes. (A) (i ) Inverted phase-contrast image of a wild type axoneme. The orange curve represents the tracked centerline. The points depict the basal end (red), the distal end (green) and the center position (black) of the axoneme. The yellow line depicts the trajectory of the basal end. (ii ) Same image as in Ai, magnified around the center region. The tangent angle ψ(s, t) is defined with respect to the lab frame. (iii ) Tangent angle of three different arc length positions (depicted in Ai ) as a function of time. The linearly increasing tangent angle corresponds to the rotation of the axoneme during swimming. Mean uncertainty for 1000 adjacent frames of the tracked shape δr· y in (iv ) (the δr· x error gives the same result) and tangent angle δψ in (v ). (B) is analogous to (A) but for mbo2 mutant axonemes. Experimental procedure Preparation, reactivation and imaging of axonemes In this study, axonemes from Chlamydomonas reinhardtii wild type cells (CC-125 wild type mt+ 137c, R.P. Levine via N.W. Gillham, 1968) and mutant cells that move backwards only (CC-2377 mbo2 mt-, David Luck, Rockefeller University, May 1989) were purified and reactivated. The procedures described in the following are detailed in [26]. Chemicals were purchased from Sigma Aldrich, MO if not stated otherwise. In brief, cells were grown in TAP+P medium under conditions of illumination (2x75 W, fluorescent bulb) and air bubbling at 24 ◦C over the course of 2 days, to a final density of 106 cells/ml. Flagella were isolated using dibucaine, then purified on a 25% sucrose cushion and demembranated in HMDEK (30 mM HEPES-KOH, 5 mM MgSO4, 1 mM 6 (i)10 µm500 nm500 nmB10 µm(i)(ii)(i)(ii)0.020.060.1−20246Time, t (s)Tangent angle (rad)0.020.060.1−20246Time, t (s)Tangent angle (rad) (µrad) (nm)Arclength, s (µm)0510010200204060 (µrad)(iii)(iii)(iv)(v)(iv)(v)Awild typembo2 mutant051005100510010200204060Arclength, s (µm)Arclength, s (µm)Arclength, s (µm) (nm) DTT, 1 mM EGTA, 50 mM potassium acetate, pH 7.4) augmented with 1% (v/v) Igpal and 0.2 mM Pefabloc SC. The membrane-free axonemes were resuspended in HMDEK plus 1% (w/v) polyethylene glycol (molecular weight 20 kDa), 30% sucrose, 0.2 mM Pefabloc and stored at −80 ◦C. Prior to reactivation, axonemes were thawed at room temperature, then kept on ice. Thawed axonemes were used for up to 2 hours. Reactivation was performed in flow chambers of depth 100 µm, built from easy- cleaned glass and double-sided sticky tape. Thawed axonemes were diluted in HMDEKP reactivation buffer containing 1 mM ATP and a ATP-regeneration system (5 units/ ml creatine kinase, 6 mM creatine phosphate) used to maintain the ATP concentration. The axoneme dilution was infused into a glass chamber, that was blocked using casein-solution (solution of casein from bovine milk, 2 mg/mL, for 10 minutes) and then sealed with vacuum grease. Prior to imaging, the sample was equilibrated on the microscope for 5 minutes and data was collected for a maximum time of 20 minutes. The reactivated axonemes were imaged by either phase constrast microscopy (wild type axonemes) or darkfield microscopy (mbo2 axonemes). Phase contrast microscopy was set up on an inverted Zeiss Axiovert S100-TV microscope using a Zeiss 63× Plan- Apochromat NA 1.4 Phase3 oil lens in combination with a 1.6× tube lens and a Zeiss oil condenser (NA 1.4). Data was acquired using a EoSens 3CL Cmos highspeed camera. The effective pixel size was 139 nm/pixel. Darkfield microscopy was set up on an inverted Zeiss Axiovert 200 microscope using a Zeiss 100× Plan-Neofluar NA iris 0.7-1.4 oil lens in combination with an 1.2× tube lens and a Zeiss oil Darkfield (NA 1.4). Data was acquired using a pco dmaxS highspeed camera. In both cases the illumination was performed using a Sola light engine with a 455 LP filter. Movies of up to 3000 frames were recorded at a frame rate of 1000 fps. The sample temperature was kept constant at 24◦C using an objective heater (Chromaphor). High precision tracking of isolated axonemes To localize the position of the axoneme shapes in each frame of the recorded movie with nm precision, the Matlab based software tool FIESTA was used [27]. Prior to tracking, movies were background subtracted to remove static inhomogeneities arising from uneven illumination and dirt particles. The background image contained the mean intensity in each pixel calculated over the entire movie. This procedure increased the signal to noise ratio by a factor of 3. Phase contrast images were then inverted; darkfield images were tracked directly. The tracking algorithm FIESTA uses manual thresholding to determine the filament skeleton, which is then divided into square segments. During tracking, the filament position in each segment is determined independently. For tracking, a segment size of 733 nm (approximately 5x5 pixels) was used, corresponding to the following program settings: a full width at half maximum of 750 nm, and a "reduced box size for tracking especially curved filaments" of 30 %. Along the arc length of each filament, 20 equally distributed segments were fitted using two dimensional Gaussian functions. Two examples of spline fitted shapes are presented in Figure 2 Ai and Bi superimposed on the tracked image. The mean localization uncertainty of the center position of each of these segments was about 5 nm (see Figure 2 Aiv and Biv ). For localization of the ends, the program uses a different fitting function, resulting in an increased uncertainty. 7 Results Quantification of the beat of wild type and mbo2 cilia To test the different mechanisms of beat regulation we performed high precision mea- surements of the flagellar waveform, see Experimental Procedures and Fig. 2. Using the tracking software developed in [27] we calculated the temporal trajectories of 20 points along the arc length of the axoneme. The uncertainty of the shape in the xy space was ∼ 5 nm and the uncertainty in the tangent angle was 20 µrad (see panels iv and v in Fig. 2). The latter corresponds to a sliding displacement between adjacent doublet microtubules of only 0.06 nm. Figure 3. Fourier decomposition of the beat. Panel A shows the Fourier decomposition of the waveform of wild type axonemes, panel B the decomposition of the mbo2 axoneme waveform. (A) (i ) Power spectrum of the tangent angle averaged over arc length. The fundamental mode (n = 1) and three higher harmonics (n = 2, 3, 4) are labeled. (ii ) Angular representation of the static (n = 0) mode as a function of arc length. The constant slope indicates an arc length independent static curvature ψ0 = C0. (iii ) The amplitude and phase (argument) of the fundamental mode are shown in iii and iv , respectively. The linear decrease in phase indicates steady wave propagation. The data of a representative axoneme is highlighted in the panels ii -- iv, with error bars indicating the standard error of the mean calculated by hexadecimation. (B) Equivalent plots to (A) for mbo2 axonemes. (C) Beat shapes of one representative beat cycle of the wild type axoneme highlighted in panel A (left panel, data) and shapes reconstructed from the superposition of the static and fundamental modes, neglecting all higher harmonics. The progression of shapes through the beat cycle is represented by the rainbow color code (see inset). (D) Same as (C) for an mbo2 mutant axoneme. 8 0200400Frequency (Hz)10-610-410-2Arc-length (μm)0510-3-2-100.00.40.81.2-6-40-2Arc-length (μm)0510Arc-length (μm)0510(i)(ii)(iii)(iv)0200400Frequency (Hz)10-610-410-2Arc-length (μm)0510-3-2-10.00.40.81.2-6-40-2Arc-length (μm)0510Arc-length (μm)05100datareconstructiondatareconstructionABCD(i)(ii)(iii)(iv)wild typembo2 mutantwild typembo2 mutant Because the beat of Chlamydomonas is periodic in time, it is convenient to decompose the tangent angle ψ(s, t) into Fourier modes ψn, see Eq. 2. Before doing so, we note that wild type Chlamydomonas axonemes swim in circles at a slow angular rotation speed ωrot ∼ 30 rad/s, see i and iii in Fig. 2A. While the effect of this rotation is small for a single beat it becomes large for a long time series. Before performing the Fourier decomposition we subtracted ωrott from the tangent angle ψ(s, t) 1. The power spectrum of the tangent angle (averaged over the flagellar length) shows clear peaks at harmonics of its fundamental frequency, see Fig. 3Ai. Note that the peak of the fundamental mode n = 1 accounts for 90% of the total power spectrum, and so we neglect the higher harmonics n = 2, 3, 4, . . . for reconstructing the flagellar shape. The amplitude of the static mode (n = 0) and the amplitude and phase of the fundamental mode (n = 1) are shown in Fig. 3ii -- iv. The main difference between the wild type and mutant axonemes comes from the static angular mode ψ0. For wild type axonemes, ψ0 decreases with an approximately constant slope over arc length. This corresponds to a constant static curvature ∼ 0.25 rad/µm, and indicates that the static mode of the shape is a circular arc of radius ∼ 4 µm. The static curvature of wild type axonemes leads to a highly asymmetric waveform. In contrast, mbo2 mutant axonemes have a very small static mode with a curvature ∼ 0.025 rad/µm, and the resulting beat is approximately symmetric. In comparison to the large difference in the static mode between wild type and mutant axonemes, the fundamental modes are similar. The amplitude of ψ1 is roughly constant and has a characteristic dip in the middle. The argument of ψ1, which determines the phase profile of the wave, decreases at a roughly constant rate in both cases. Since the total decay is about −2π, the wavelength of the beat is approximately equal to the length of the axoneme. In summary, both wild type and mutant cells have an approximately sinusoidal dynamic beat whose amplitudes drop in the middle of the axoneme and whose phases decrease monotonically, consistent with a beat wavelength of about 13 µm. Motor regulation in the axoneme: experiments vs theory The response of the motor force to strains and stresses of the axoneme described by Eq. 6 allows for regulation via sliding, curvature and normal forces. While it is possible that all three mechanisms of motor control are involved in regulation of the axonemal beat, we now show each individual mechanism alone is capable of producing dynamic bending patterns. Because the axonemal beat is dominated by its static mode (n = 0) and fundamental dynamic mode (n = 1), see Fig. 3, we constrain our physical model to one dynamic mode, which corresponds to the critical point of an oscillatory Hopf bifurcation [7, 25]. Within our theoretical description, sliding control corresponds to the case in which β = 0 and γ = 0, and the motor force only responds to sliding changes through χ(ω) = χ(cid:48) + iχ(cid:48)(cid:48), with single and double primes denoting real and imaginary parts respectively and i the imaginary unit. Additionally, because the active response must dominate for oscillations to occur, we have χ(cid:48) < 0 and χ(cid:48)(cid:48) < 0 [6, 25]. In curvature control, the motors have a passive response to sliding (corresponding to the slope of their force-velocity curves), so that χ(cid:48) > 0; the motors are actively regulated by curvature, thus β(cid:48)(cid:48) < 0; and they are not regulated by normal force, that is γ = 0. Finally, in normal force control there is a passive response to sliding, χ(cid:48) > 0, an active response to normal forces, with γ(cid:48) > 0 and γ(cid:48)(cid:48) < 0, and no response to curvature, β = 0. 2 For wild type cells, we compared the modes obtained from the three motor models with the observed beats (see Appendix for details on fitting procedure and parameters 1For simplicity we use the same notation for ψ(s, t) and ψ(s, t) − ωrott. 2For backward traveling waves the signs of β(cid:48), β(cid:48)(cid:48) and γ(cid:48), γ(cid:48)(cid:48) change, but not those of χ(cid:48), χ(cid:48)(cid:48). 9 Figure 4. Comparison of theoretical and experimental beating patterns. (A) Comparison of the theoretical (lines) and experimental (dots) beating patterns of a typical wild type axoneme. The real and imaginary part of the first mode of the tangent angle ψ1(s) is plotted for beats resulting from sliding control, curvature control, and normal force control. (B) Analogous to (A) for mbo2. Note that here also curvature control and normal force control provide good agreement, but not sliding control. (C) and (D) Theoretical and experimental shape reconstruction in position space for the wild type and mbo2 beats under curvature control. used). The result of a typical fit is shown in Fig. 4A, where the only two free parameters were the basal stiffness kb and viscosity ξb. As we can see, the predicted real and imaginary part of ψ1(s) agree well with the data for a wild type axoneme in the cases of curvature control and normal-force control, but not for sliding control. In fact, in the beats predicted by sliding control, the real and imaginary parts are in anti-phase. This results in a standing wave with no bend propagation. The beating pattern obtained by curvature control is compared to the experimental reconstruction in Fig. 4C. The good agreement reinforces the conclusion from Figure 4A that the curvature control model accords with the experimental data for wild type cells. Similar good agreement for wild type cells was found with the normal-force model. Table 1 summarizes average parameters resulting from the fits of 9 different axonemes. We compared theory and experiments for the symmetric beat of the mutant mbo2, where the static curvature is reduced by at least one order of magnitude compared to wild type beats, see Fig. 3. In this case the results were similar to those of wild type, see Fig. 4B: sliding control cannot produce bend propagation, while curvature and normal force control are in good agreement with the experimental data. The parameters obtained from the fit of mbo2 beats are given in Table 2. 10 Sliding controlCurvature controlNormal force control0.00.20.40.60.81.0-1.0-0.50.00.51.0Arc-length, s/LR20.52R20.95R20.950.20.40.60.81.0-1.0-0.50.00.51.0Arc-length, s/L0.20.40.60.81.0-1.0-0.50.00.51.0Arc-length, s/L0.00.0Sliding controlCurvature controlNormal force controlR20.67R20.97R20.970.20.40.60.81.0-1.0-0.50.00.51.0Arc-length, s/L0.00.20.40.60.81.0-1.0-0.50.00.51.0Arc-length, s/L0.00.20.40.60.81.0-1.0-0.50.00.51.0Arc-length, s/LRe(Ψ1)&Im(Ψ1)0.0wild typembo2 mutanttheoryexperimenttheoryexperimentRe(Ψ1)&Im(Ψ1)Re(Ψ1)&Im(Ψ1)Re(Ψ1)&Im(Ψ1)Re(Ψ1)&Im(Ψ1)Re(Ψ1)&Im(Ψ1)CDBAwild typembo2 mutant====== Table 1. Parameters for beat generation in wild type axonemes. 0 b (nN· µm−1) Sliding control 49 ± 4 Curvature control Normal force control 95 ± 1 24 ± 4, 0 0,−7.1 ± 0.5 0 51 ± 15, 2.6 ± 0.9 −33 ± 11 19.5 ± 4.7 R2 (%) χ(cid:48), χ(cid:48)(cid:48) (nN· µm−2) −14.7 ± 3.6,−1.31 ± 0.16 β(cid:48), β(cid:48)(cid:48) (nN) γ(cid:48), γ(cid:48)(cid:48) χ(cid:48) b, χ(cid:48)(cid:48) ∆b,0 (nm) ∆b,1 (nm) Curvature control and normal force control result in high R2 values. Sliding control, unable to produce bend propagation (see Fig. 4), produces a low R2. The values reported are mean and standard deviation calculated with 9 axonemes (when the standard deviation was larger than the mean it was replaced by the mean itself). The average static curvature is C0 = −0.232 ± 0.009µm−1. Note that, for curvature control beats, using the value of β here reported and an estimate curvature of 0.1 rad/µm results in a sliding force of ∼ 700 pN/µm. Since the motor density of an active half of the axoneme is ∼ 500 µm−1 the individual motor force is about ∼ 1 pN. For the case of normal force control the sliding force generated by the motors is of the order of the normal force that they experience, since γ ≈ 2. 95 ± 1 16.0 ± 4.4, 0 0 0.123 ± 0.052, 1.96 ± 0.21 96.3 ± 139.0, 15.2· 103 ± 40.9· 103 −131 ± 178 10.7 ± 9.7 0 51.3 ± 0.9, 3.35· 103 ± 7.16· 103 −37 ± 37 21 ± 17 Table 2. Parameters for beat generation in mbo2 mutant axonemes. 0 b (nN· µm−1) Sliding control 72 ± 5 Curvature control Normal force control 95 ± 1 21 ± 3, 0 0,−7.1 ± 0.5 0 2.5 ± 0.2, 2.8 ± 2.1 −66 ± 40 54 ± 10 R2 (%) χ(cid:48), χ(cid:48)(cid:48) (nN· µm−2) −18.2 ± 1.2,−0.52 ± 0.12 β(cid:48), β(cid:48)(cid:48) (pN) γ(cid:48), γ(cid:48)(cid:48) b, χ(cid:48)(cid:48) χ(cid:48) ∆b,0 (nm) ∆b,1 (nm) As for wild type beats curvature control and normal force control provide very good fits, and sliding control does not. Values indicated are averages and standard deviations for 9 axonemes (when the standard deviation was larger than the mean it was replaced by the mean itself). The static curvature was C0 = −0.0276 ± 0.005µm−1. Note that the values of γ in normal force control are very spread and different relative to those obtained for wild type fits. In fact, in one case we obtained γ ≈ 80, indicating that motors must amplify the normal force they sense by almost two orders of magnitude. The values for curvature control are very similar to those of wild type fits. . 96 ± 1 12.7 ± 5.2, 0 0 1.52 ± 1.52, 32 ± 25 2.8 ± 0.3, 16.4 ± 7.2 −58 ± 36 69 ± 25 0 37 ± 17, 1.98 ± 0.87 −5.7 ± 5.1 46 ± 9 Regulation of the beat by sliding There are two related reasons why the sliding control mechanism provides poor fits to the observed beating patterns. First, under sliding control the equations describing the flagellar beat become symmetric with respect to a change in sign of the arc length. As a consequence, the only possible solutions are standing waves. This can be most easily seen by using the plane-wave approximation, ψ1 = exp(−2πis/λ) with λ real, for a symmetric beat (C0 = 0) regulated by sliding (β = 0 and γ = 0). In that case, Eq. 4 becomes an equation for the wave-length i¯ω = −(2πL/λ)4 − (2πL/λ)2 ¯χ , (7) 11 where ¯χ = a2 0L2χ/κ is the dimensionless complex sliding response coefficient. Note that this equation is symmetric with respect to the change λ → −λ, and thus admits simulta- neously waves traveling in both directions. In the absence of boundary asymmetries, these opposing waves interfere to form a standing wave. Thus to the extent that the asymmetry of the boundary can be neglected, the sliding control mechanism can not account for the bend propagation. Figure 5. The role of length in sliding regulated beats. (A) Wave speed versus relative length of the first two unstable waveforms of a freely swimming axoneme regulated by sliding. For short lengths, in the range of Chlamydomonas, the modes loose directionality and become standing waves. Long axonemes have directional waves that can travel either forward, as in waveform 1, or backwards, as in waveform 2. (B) Two examples of waveform 1 for long (top) and short (bottom) axonemes. Note that the short axoneme shows a standing wave. (C) beating patterns of waveform 2 for a long (top) and short (bottom) axonemes. In B and C arrows denote direction of wave propagation. The second reason for the poor fits of the sliding control mechanism is that short flagella can not exhibit bend propagation, as already noted in [7]. This is true even in the presence of a boundary asymmetry (from the basal compliance). To understand this, we note that the dimensionless parameter ¯ω can be written as (cid:18) L (cid:19)4 (cid:96) (cid:18) κ (cid:19)1/4 ξnω ¯ω = with (cid:96) = , (8) v =(cid:82) L where (cid:96) is the characteristic length at which oscillations decay in a boundary driven axoneme [28]. For the case of flagella short with respect to (cid:96), we can approximate ¯ω ≈ 0 √−χ and λ− = −λ+, where one in Eq. 7. The two allowed wavelengths are λ+ = 2πL/ can show that ¯χ = −π2/4 in order to satisfy the boundary condition of no torques on the distal end [7]. These two modes correspond again to opposing waves, and imposing no torques on the basal end it can be shown that they must have equal amplitudes. The result is again a standing wave, irrespective of the basal asymmetry. To verify this argument we studied the speed of bend propagation, which is defined as 0 ψ12∂s arg ψ1ds, for two different flagellar lengths L, see Fig. 5A. For axonemes with lengths L similar or smaller than (cid:96), the resulting beats exhibit no bend propagation, Figs. 5 B and C bottom. For axonemes significantly longer than (cid:96) wave propagation can be strong, Figs. 5 B and C top. While in bull sperm we have L/(cid:96) ≈ 8, which is enough to produce wave propagation [8]; for Chlamydomonas we have L/(cid:96) ≈ 2, which results in almost no bend propagation. Indeed, we estimated the wave speed of the sliding control waveform in Fig. 4 to be two orders of magnitude smaller than the observed value. Regulation of the beat by normal forces Despite the good fits obtained for wild type and mbo2 axonemes, there are two obser- vations that argue against normal force control. The first is that unlike the curvature 12 waveform 2waveform 1BCwaveform 1waveform 2ChlamydomonasBull sperm24681012A-2wave speed, v-1012 response coefficient β, the mean value of the normal-force response coefficient γ is much greater in mutants cells than in wild type cells (Table 2). The reason is that in mbo2 the small static curvature results in a small normal force, which requires a correspondingly larger response coefficient γ. In other words, to obtain agreement between the observed and theoretical waveforms, the sensitivity of the motors to normal force needs to be much greater in the mbo2 mutant than in the wild type. Such a big difference in the response coefficient is not expected, given that the dynamic components of the mutant and wild type waveforms are very similar (compare Fig. 3C and 3G). Figure 6. Scaling of response coefficients with asymmetry. (A) Since the normal force f⊥,1 is proportional to the static curvature C0 but the fundamental mode stays unchanged, the normal force response coefficient γ is inversely proportional to the curvature. In red values for mbo2 and in green for wild type. (B) The curvature control response coefficient β(cid:48)(cid:48) is independent of the asymmetry, and stays constant even for a change in asymmetry of several orders of magnitude. The second argument against the normal force control is that γ varies greatly from cell to cell for the case of mbo2. This is due to the observation that, while the static curvature is small in mbo2 axonemes, it is variable. This leads to a large variability in γ from axoneme to axoneme, which correlates strongly with the inverse of the static curvature (see Fig. 6A), i.e. γ ∝ C0−1. Thus, even though the amplitude and phase of the first mode in mbo2 axonemes is very similar from axoneme to axoneme (Fig. 3B (iii) and (iv)), the response coefficients vary widely in amplitude. The reason for this correlation between γ and C0 is that, according to Eq. 5, the dynamic component of the normal force is linearly proportional to the asymmetry C0. This means that to preserve a similar fundamental dynamic mode, axonemes with a smaller static asymmetry C0 require motors to have a higher response coefficient γ. By contrast, the curvature control response coefficient is highly consistent for wild type and mutant axonemes. In summary, despite the similarities in the dynamics of the beats of wild type and mbo2 axonemes, the normal force model requires very different values for the response coefficient γ (and also the basal response coefficient χb) for wild type and mbo2 axonemes. Furthermore, the normal force model requires very large differences in γ from axoneme to axoneme, despite the similarity in the dynamics between axonemes. Thus, we conclude that normal force is not a plausible parameter for controlling the ciliary beat. Regulation of the beat by curvature The curvature control model provides a good fit to the experimental data for both wild type and mbo2 axonemes (Figure 4A and B, middle panel). However, the strategy of curvature regulation that we used is strikingly different from those previously studied [8, 9, 11, 19, 22]. While in previous work it was assumed that the static and dynamic response to curvature are of equal importance, here we found that when χ(cid:48)(cid:48) and β(cid:48) were unconstrained, their best fit values were not significantly different from zero. We therefore set them both to zero (see Table 1). We can understand this by writing Eq. 4 13 radius of curvaturenormal force responsecoefficient050100150200020406080mbo2wild-typeradius of curvature050100150200curvature responsecoeficcient , nN02468mbo2wild-typeAB for symmetric plane waves regulated by curvature, which, after separating real and imaginary part, becomes ¯ω + (2πL/λ)2 ¯χ(cid:48)(cid:48) − (2πL/λ)3 ¯β(cid:48) = 0 (2πL/λ)4 + (2πL/λ)2 ¯χ(cid:48) + (2πL/λ)3 ¯β(cid:48)(cid:48) = 0 , (9) where ¯β(cid:48) = a0Lβ(cid:48)/κ is the dimensionless real curvature response coefficient, ¯β(cid:48)(cid:48) = a0Lβ(cid:48)(cid:48)/κ is the complex one, and we assume a passive response to sliding, χ(cid:48) > 0 and χ(cid:48)(cid:48) > 0. These equations show that for a wave traveling forward (λ > 0), β(cid:48) > 0 produces the active force to counter viscous effects of fluid and filament sliding, while β(cid:48)(cid:48) < 0 counters the elastic forces of filament bending and sliding. For the case in which the contribution of the response to sliding is smaller than that of the fluid and filament bending, we can divide the equations above to obtain β(cid:48)/β(cid:48)(cid:48) ∼ ¯ω/(2πL/λ)4. Since for Chlamydomonas ¯ω ∼ 24 and λ ∼ L, we find that β(cid:48)/β(cid:48)(cid:48) ∼ 10−2. In other words, for Chlamydomonas, elastic forces dominate over viscous forces, and thus β(cid:48)(cid:48) (cid:29) β(cid:48), as we obtained from the fits. Furthermore, in the limit of ¯ω = 0 (short lengths, low viscosity), we can set χ(cid:48)(cid:48) = β(cid:48) = 0 and still obtain plane waves. The opposite however is not true: for β(cid:48)(cid:48) = 0 the balance of elastic forces can not be satisfied. The molecular implications of this finding will be expanded upon in the Discussion section. Figure 7. Phase space of curvature control. (A) Heat map of the mean square distance R2 between the theoretical and a reference experimental beat as a function of the sliding response coefficient χ(cid:48) and the curvature response coefficient β(cid:48)(cid:48). The ellipsoid delimits the region with R2 = 0.90. Black lines delimit the region with a passive base. Moving along the long axis (green circles) affects the amplitude dip in the midpoint of the axonemem, see left panel in (B). Moving along the small axis towards the region of active base results in waveforms with a large amplitude at the base (blue and red circles), see central and right panels in B. The axis in A are normalized by the reference fit, such that (χ(cid:48) = 1, β(cid:48)(cid:48) = 1) corresponds to the highest value of R2. b and χ(cid:48)(cid:48) We therefore set χ(cid:48)(cid:48) = β(cid:48) = 0 in the subsequent analysis. By doing so, we significantly simplify the model, because the number of free parameters is now only two, χ(cid:48) and β(cid:48)(cid:48). Note that the two parameters, χ(cid:48) b, which characterize the stiffness and viscosity at the base respectively, are determined once χ(cid:48) and β(cid:48)(cid:48) are specified as we are looking for oscillating solutions to the boundary value problem (see Appendix). Thus, the curvature control model is specified by just two free parameters, χ(cid:48) and β(cid:48)(cid:48), which are specified by the sliding elasticity between doublet microtubules and the rate of change of axonemal curvature. The average values of χ(cid:48) and β(cid:48)(cid:48) varied little between wild type and mbo2 mutant axonemes (compare the third column of Table 1 with that of Table 2). This accords with the observation that there is little difference in the dynamical properties of the beat between wild type and mbo2 axonemes. Furthermore, the standard deviation of χ(cid:48) and β(cid:48)(cid:48) are small, indicating that there is little variation from axoneme to axoneme. Thus, the tight distribution of values of the parameters in the model reflects the similarity in the observed shapes in different axonemes. In other words, χ(cid:48) and β(cid:48)(cid:48) are well constrained by the experimental data. 14 To understand what aspects of the experimental data specify these two parameters, we performed a sensitivity analysis on χ(cid:48) and β(cid:48)(cid:48). In Fig. 7A we show a density map of the mean square distance R2 between the theoretical waveforms and a reference experimental beating pattern as a function of χ(cid:48) and β(cid:48)(cid:48). A red ellipse delimits the region with R2 > 0.90, which very closely coincides with the region where χ(cid:48) b are both positive, delimited by black lines. This is very important because negative values of the basal parameters imply an active process at the base. Such an active process would drive a whiplike motion of the cilium as discussed in [12]. Evidently, the observed shapes of the beats rule out such a whiplike motion. To investigate how the beat pattern is affected by variations in χ(cid:48) and β(cid:48)(cid:48), as well as the existence of active processes in the base, we systematically varied χ(cid:48) and β(cid:48)(cid:48) parallel and perpendicular to the long axis of the ellipse. Moving parallel affects the amplitude of the beat, with the middle-dip becoming more or less prominent, see green shapes in Fig. 7B. Moving perpendicular in the region of active base indeed results in whiplike beats, with a large amplitude at the base, see blue and red shapes in Figs. 7B. b and χ(cid:48)(cid:48) Figure 8. Axonemal variability in phase space. (A) Circles represent values obtained from fits for each of the axonemes, green corresponds to wild type and red to mbo2. In the background we have the same heat map as in Fig. 7. Note that mbo2 points lie away from wild type circles in the direction of the short axis of the ellipse. All values are normalized by those of the reference fit used also to normalize the heat map axis. (B) The distance of the circles to the long axis of all fits shows a clear correlation with the axonemal length. Note also that mbo2 axonemes are systematically shorter than wild type axonemes. (C) The basal response coefficient also correlates with the length, resulting in a high value for wild type axonemes, which are longer. The values are normalized by the value for a reference axoneme. To better understand the cell to cell variability we placed all the axonemes recorded in the (χ(cid:48), β(cid:48)(cid:48)) space, see Fig. 8A. Points scatter mainly along the long axis of the ellipse, where there is a large region of small shape variations. Importantly, we consistently see a shift perpendicular to the long axis between the wild type and mbo2 mutant axonemes. This variation mainly comes from the difference in length between wild type and mutant axonemes, as can be seen in Fig. 8B. The implication of this length variation is a stiffening of the base, see Fig. 8C, which explains the variability in basal compliance between mbo2 and wild type axonemes for curvature control in Tables 1 and 2. Discussion In this work we imaged isolated axonemes of Chlamydomonas with high spatial and temporal resolution. We decomposed the beating patterns into Fourier modes and compared the fundamental mode, which is the dominant dynamic mode, with theoretical predictions of the three motor control mechanisms illustrated in Fig. 1. The sliding control model provided a poor fit to the experimental data. We argued that the reason for this is that sliding control cannot produce wave propagation for axonemes as short as those of Chlamydomonas, see Fig. 5. While the normal force model provided good fits to 15 the experimental data, it relies on the presence of static asymmetry [22, 29], which varies greatly between the mbo2 and wild type axonemes. As a result of this large difference in static curvature, the control parameters in this model had to be varied over a large range to fit the data from the different axonemes, see Fig. 6. Because the waveforms of mbo2 and wild type axonemes have very similar dynamic characteristics, such variation in the control parameter seems implausible. Finally, the curvature control model provided a good fit to the experimental data with similar parameters for mbo2 and wild type axonemes. Thus, we conclude that only the two-dimensional curvature-control model is fully consistent with our experimental data. A potential caveat of the model used here is that it is two-dimensional. Importantly, in order to simplify the geometry, the model only contains one pair of doublet microtubules. In the three-dimensional axoneme, there are pairs of doublets on opposite sides of the axoneme which, due to the approximate rotational symmetry of the structure, are bent in opposite directions when their associated dyneins are activated. A key feature of dynein coordination is that motors on either side of the axoneme are antagonistic (i.e. in a "tug-of-war"), such that when dyneins on either side are active they bend the axoneme in opposite directions [16]. Both sliding control and curvature control ensure that the tug-of-war is unstable such that if dyneins on one side begin to dominate, then they completely dominate in a "winner-takes-all" scenario. To capture this idea of reciprocal inhibition by opposing dyneins in the two-dimensional model, dyneins are anchored with opposite orientations to each of the doublets in the pair [18]. While this captures the essential features of the sliding control and curvature control models, it oversimplifies the normal force model, because in the three-dimensional axoneme there are radial and transverse forces acting on the doublets as the axoneme bends. Yet the two-dimensional model does not distinguish between them. To bridge this gap, in other work [30] we use a full three-dimensional description of the axoneme to calculate the radial and transverse stresses. The three-dimensional model shows that even when there is a static curvature (without twist), normal (transverse) forces are not antagonistic across the centerline and therefore cannot serve as a control parameter for motors. Relation with past work Earlier results showed that sliding control can account for the beating patterns of sperm [5, 8, 14]. This result is not inconsistent with our results because the bull sperm axoneme is approximately five times longer than the Chlamydomonas axoneme and we have shown that sliding control can work for long axonemes, while for short axonemes it produces no bend propagation, see also [31]. Thus, it is possible that different control mechanisms operate in different cilia and flagella, with sliding control being used in longer axonemes and curvature control being used in shorter ones. However we do note that curvature control models can account for the bull sperm data [8] as well as data from other sperm [29, 32, 33], so there is no strong morphological evidence favoring either sliding or curvature control in sperm. Other studies have shown that the normal force model gives beating patterns that resemble those of sperm [13, 29, 34]. These models, like ours, rely on there being an asymmetry. What we have shown is that the control parameter depends critically on this asymmetry while the similarity in the fundamental dynamic modes of mbo2 and wild type suggests that the parameters should be similar. Thus, the curvature control model, unlike the other two models, robustly describes symmetric and asymmetric beats in short and long axonemes, and could serve as a "universal" regulator of flagellar mechanics. 16 Figure 9. Phase delays and active force distribution on the axoneme. (A) Polar representation of the phase of sliding ∆, curvature ψ, normal force f⊥ and the sliding force f for a beating axoneme regulated by curvature. For a plane wave we have ψ1 = exp i(ωt − 2πs/λ). A time derivative thus adds a counter-clockwise π/2 phase, while a spatial derivative adds π/2 clockwise. Deviations from this come from the solution deviating from a plane wave due to the boundary conditions. In dashed red we show Machin's prediction for an optimum flagellum. (B) Time evolution of the quantities in A. (C) Beating axoneme with two shapes highlighted during the breast and recovery strokes. The black arrows represent direction of motion, and the circular arrows represent the local sliding force. (D) Sliding force density over arc length for the two shapes highlighted in C. Dynamic curvature control as a mechanism for motor regulation One interesting feature of our curvature control model is that in order to describe the observed beating patterns the motor force depends only on the time derivative of the curvature. This follows from the fact that the curvature response function β(ω) has no real part, see tables, and thus vanishes at zero frequency, β(ω = 0) = 0. Such a model is fundamentally different from the current views of curvature control, in which motors are thought to respond to curvature [10 -- 12, 32] and not to its time derivative. While motors can respond to time derivatives of sliding displacement through their force-velocity relation, it is hard to understand how a similar mechanism could apply to curvature. A more plausible mechanism giving rise to a response to the time derivative of curvature is an adaptation system analogous to that of sensory systems, like the signaling pathway of bacterial chemotaxis [35 -- 38]. In an adaptation mechanism motor activity would be "remembered", and the average activity over the past times would in turn down-regulate the activity of the motor on a long time-scale. Such regulation could occur, for example, via phosphorylation sites in the dynein regulatory complex or the radial spokes [39 -- 42]. Just as methylation of the chemoreceptors of bacteria modifies their ligand affinity, phosphorylation of regulatory elements within the axoneme could modify the motor sensitivity to curvature. Independence of static and dynamic waveform components Our dynamic curvature control model adds to the view that dynamic and static com- ponents of the beat are regulated independently. The problem with models in which dynein activity is regulated by the instantaneous value of the curvature [10 -- 12, 22] is 17 PhaseABTime, Breast strokeRecovery strokeArc length, s-220CD0L/2LForce, ( ) Arc length, s0L/2LBreast strokeBreast strokeRecovery stroke that both the static and dynamic components of the beat would contribute to regulation and hence the dynamic component of the waveform would be highly dependent on the static component [43]. This potential problem was noted by Eshel and Brokaw in [21]. Our dynamic curvature control model provides a solution to this problem because static curvature is "adapted" away. We now bring together several lines of evidence supporting the notion that the static and dynamic modes are separable in their origin and in their affect on the beat 1. The waveform of Chlamydomonas can be mathematically decomposed into a bending wave superimposed on a static asymmetry (a circular arc), see [15, 21]. 2. Dynamic and static components of the beat can exist independently of each other. This is evidenced by the existence of bent non-motile cilia at low ATP concentrations on the one hand, as well as symmetrically beating mutants on the other [15]. 3. The waveform of mbo2 has a similar fundamental dynamic mode as that of wild type, compare Fig. 3 A (iii) and B (iii). However, the static mode is absent in the former, compare Fig. 3 A (ii) and B (ii). The same also holds for the two beating modes of the uniflagellar mutant [21]. Thus, altering the static mode of Chlamydomonas has little effect on the dynamic mode. 4. The dynamic motor response coefficients are largely independent of the asymmetry, and very similar for mbo2 and wild type axonemes, see Fig. 8 A. If the dynamic and static modes are indeed independently controlled, the dynamic motor response is robust to changes in the asymmetry. This has important biological implications: power generation (the beat) and steering (the asymmetry) can be inde- pendently controlled so that the swimming direction can be adjusted without having to alter the motor properties. Phase relations during the beat The dynamics of the axoneme can be simplified using a phase plot, see Fig. 9A. In this diagram, each arrow represents the normalized complex value of the corresponding quantity at the midpoint of the axoneme (s = L/2). Over time, the arrows rotate counter-clockwise at a homogeneous speed, preserving their phase relations. In the plane wave limit, an arc length derivative adds a delay of π/2, which is why the curvature lags almost one quarter behind the sliding. The motor force is itself delayed one quarter of a cycle with respect to curvature, which makes it in anti-phase with the sliding. Finally, since C0 < 0, from Eq. 5 we see that the normal force is in anti-phase with the curvature, see Fig. 9B for a time trace. Importantly these phase relations deviate from the prediction made by Machin in [12] for the most energy efficient flagellum (note that his bending moment B is ∼ −F , and his displacement y is ∼(cid:82) ∆ds). In fact, using the same plane wave arguments as in [12], we already showed after Eq. 9 that the elastic contribution to the sliding response dominates the viscous component by almost two orders of magnitude, which results in the one quarter delay with respect to curvature. Thus, the Chlamydomonas flagella is not optimal under Machin's assumptions. 18 Appendices Non-linear dynamics of the axoneme The equations that describe the dynamics of the axoneme are obtained by balancing mechanical and fluid forces. We proceed using a variational approach similar to that in [7, 8, 22], and introduce the work functional U given by (cid:21) (cid:90) L (cid:20) κ 2 0 U = ψ2 + f ∆ + f⊥(a − a0) + Λ 2 (r2 − 1) ds + kb 2 ∆2 b , (10) where κ is the bending rigidity, f ∆ the work performed by the motors, and kb the stiffness of cross-linkers at the base. The normal force f⊥ is a Lagrange multiplier that ensures that a = a0, see [22, 29, 43] for the more general case of variable a. Similarly, Λ is a multiplier that ensures the incompressibility constraint r2 = 1, and is related to the tension through τ = Λ + κ ψ2 − a0F ψ, where F = −(cid:82) L lating it requires computing δ ψ. From the relation r(s) = r0+(cid:82) s (cid:48) s f (s(cid:48))ds The mechanical force that the axoneme exerts on the fluid is given by δU/δr, and calcu- 0 (cos(ψ(s(cid:48))), sin(ψ(s(cid:48))))ds(cid:48), where r0 is the position of the base, it follows that ψ = n· t and δ ψ = n· δr. Using this, we arrive at δU/δr = ∂s [7, 43]. Similarly, the net sliding force exerted at the base is δU/δ∆b = −F (0) + kb∆b [22]. To obtain the dynamics of the axoneme we balance these mechanical forces by the fluid friction Π· ∂tr and the basal friction ξb∂t∆b, which results in ∂tr = −(ξ−1 ∂t∆b = −ξ−1 t tt)· ∂s n nn + ξ−1 b (kb∆b − F (0)) (κ ψ − a0f )n − τ t (κ ψ − a0f )n − τ t (cid:104) , (cid:105) (11) (12) (cid:104) (cid:105) [7]. We can also calculate a dynamic equation for the tangent angle using that ∂t r = n∂tψ, which results in ∂tψ = ξ−1 n (−κ .... ψ + a0 f + ψ τ + τ ψ) + ξ−1 t ψ(κ ψ ψ − a0f ψ + τ ) . (13) This equation contains no information about the trajectory of the basal point r0(t), which can be determined from the condition that the total force on the cilium vanishes [23,24,43]. The tension τ and normal force f⊥ are obtained by imposing the corresponding constraints. For the case of the tension we take the time derivative of r2 = 1. This gives t· ∂t r = 0, where we can replace the dynamic equation for r. For the normal force f⊥ we use the force balance δG/δa = 0 [7, 22]. The resulting constraint equations are τ − ψ2τ = − ψ(κ ... ψ − a0 f ) + ξn ξt ∂s[ ψ0(a0f − κ ψ)] ξn ξt f⊥ = F ψ . (14) Finally, to completely characterize the dynamic equations we need to use boundary conditions. These represent force and torque balances at the ends of the filament pair, and are obtained from the boundary terms of the variational calculation. For the case of free ends considered in this work we have s = 0 : κ ψ(0) = a0f (0) s = L : κ ψ(L) = a0f (L) , , τ (0) = 0 , κ ψ(0) = a0Fb τ (L) = 0 , ψ(L) = 0 , (15) where the basal force is Fb = kb∆b + ξb∂t∆b. Together with a suitable motor model that provides the dynamics of the sliding force (see Eq. 6), the equations above allow to compute the state of the cilium over time. 19 Sign convention It is convenient to clarify the sign convention used in this paper for the geometry and the forces. The tangent angle is measured with respect to the horizontal x axis and grows counter-clockwise (the xy frame has the usual orientation, see Fig. 1A ). With this choice the mid-point of the shape shown in Fig. 2A (i) and (ii), in black, has a negative angle. Since the flagellum swims counterclockwise, the tangent angle slowly grows positive over time, Fig. 2A (iii). This applies to all flagella used in this work, which were imaged using an inverted microscope such that axonemes were seen "from behind". With this convention the mean angle has a negative slope, as shown in Fig. 3A (ii), which corresponds to C0 < 0. From Eq. 1 we then have ∆ < 0 for the simple case in which ∆b = 0. A shape with C0 < 0 and ∆b = 0 requires that the bottom filament stands out at the tip, which fixes the sliding sign: sliding is positive if the bottom filament slides towards the distal end. Thus basal and distal sliding in Fig. 1A are negative. Because dynein is a minus end directed motor, the sliding force density f is taken to be positive when the sliding is negative. This corresponds to a motor with its head (green circle in Fig. 1A) in the bottom filament moving towards the base. In other words, the force is positive when the bottom filament is tensed towards the distal end and the top one towards the basal end. According to Eq. 12 a positive static force f > 0 is opposed by a basal force Fb < 0, which results in negative sliding. Such a positive force, like the one given by Eq. 3 for C0 < 0, corresponds to a negative value of the integrated sliding force F . The static force balance κ ψ0 = a0F0 establishes then that a negative integrated force results in a negative curvature, as is the case for the Chlamydomonas axoneme. Asymmetric equation for the fundamental mode The periodic dynamics of the tangent angle can be decomposed in Fourier modes as indicated in Eq. 2. For asymmetric beating patterns in which ψ0 (cid:54)= 0, the static mode is characterized by the force balance κ ψ0 = a0F0, obtained from integrating Eq. 13 and using the boundary conditions. While the static component of the tension vanishes, the normal force has a static contribution given by f⊥,0 = κ ψ2 0/a0. The dynamics of a small amplitude oscillation dominated by the fundamental mode can be described by expanding Eqs. 13 and 14 around the static component. This results in iωξnψ1 = −.... 0τ1 = − ψ0(κ τ1 − ψ2 ξn ξt ψ1 + a0 f1 + ψ0 τ1 + ψ0τ1 + ψ0(κ ψ0 ψ1 − a0 ψ0f1 + τ1) , ξn ξt ... ψ 1 − a0 f1) + ξn ξt ∂s[ ψ0(a0f1 − κ ψ1)] , (16) where non-linear terms in the fundamental mode have been neglected. This pair of equations is the generalization of the equations for the symmetric beat [12, 25]. In them, the fundamental mode is coupled to the static mode. Expanding the expression of the normal force in Eq. 14, we obtain that its fundamental mode is given by f⊥,1 = ψ0(F1 + κ ψ1/a0). The equations above can be rendered into dimensionless form using the following rescalings: ¯s = s/L, ¯∆ = ∆/a0, ¯f = a0L2f /κ, ¯τ = L2τ /κ, ¯kb = a2 0Lkb/κ and 0Lωξb/κ. This choice results in the additional rescalings ¯f⊥ = a0L2f⊥/κ, ¯ξb = a2 0L2χ/κ, ¯β = a0Lβ/κ and ¯γ = γ, since γ is already dimensionless. For the ¯χ = a2 particular case in which the static shape has constant curvature ψ0 = C0, Eqs. 16 then reduce to the asymmetric beat equations used in the main text. Eqs. 4 together with Eq. 6 form a system of ordinary differential equations. Using boundary conditions the discrete spectrum of solutions can be obtained, see [7, 43, 44]. 20 value problem is ψ1 = (cid:80)7 While the system is of sixth order, it contains an integral term in the expression for the normal force, Eq. 5. It is thus convenient to convert the system to seventh order by taking the derivative of Eq. 5, which eliminates the integral term. Provided values for the response coefficients χ, β and γ we can then use the ansatz ψ1 = Aeks to obtain a characteristic polynomial of order seven in k. The general solution to the boundary i=1 Aiekis, where the roots ki(χ, β, γ) of the characteristic polynomial are implicit functions of the motor response coefficients. The amplitudes Ai are then determined, up to an arbitrary factor, imposing that the boundary conditions be satisfied. Determining the amplitudes will in turn result in a fixed discrete spectrum of solutions for the possible basal compliances. Conversely, if the basal compliance is provided, calculating the amplitudes will return a discrete set of solutions for the real and imaginary parts of one of the response coefficients. This discrete set are the critical modes in [7], of which here we have shown two examples in Fig. 5. Fitting procedure The fitting procedure was done as follows. Given a set of values for the response coefficients χ, β and γ, a solution ψthe was obtained in the manner described in the previous section, up to an arbitrary complex amplitude. Given this solution, the force balance χb∆b = F (0) allows to determine χb. If the value for the real or imaginary parts of χb were negative, corresponding to an active base, the solution was discarded. If they were positive, then the complex amplitude was chosen as to minimize the mean square displacement R2(ψthe, ψexp) given by (cid:80)N (cid:80)N i=1 ψexp(si) − ψthe(si)2 i=1 ψexp(si) R2 = 1 − , (17) where N = 20 and the points si were equally spaced along the axonemal length. Finally, a value of R2 < 1 was obtained. This Q function, which takes as input response coefficients, was maximized with the routine FindMinimum of Mathematica 10 using the Principal Axis method. Estimation of mechanical parameters of the axoneme The only parameters entering the problem are the effective stiffness κ and spacing a0 of the opposing filament description, and the two friction coefficients ξn and ξt. For each motor model the motor response was used to best fit the data, and the basal compliance was obtained by solving the problem as described above. A doublet has 24 protofilaments compared to 13 in a microtubule. We thus estimate the bending stiffness to be κdb ≈ 2κmt ≈ 46 pN µm2, with κmt ≈ 23 pN µm2 the microtubule stiffness measured in [45]. If we consider for simplicity an additive effect among the sliding doublets of the axoneme, we have κ = 9κdb ≈ 400 pN µm2, comparable to measurements of sea urchin sperm [46]. The radius of the axoneme is approximately 0.2 µm [47]. Considering that three doublets can be simultaneously subjected to significant active sliding during the beat we estimate a0 ≈ 0.06 µm. The friction coefficients are ξt ≈ 2πµ/(ln(2L/a0) − 1/2) and ξn ≈ 2ξt, where µ is the viscosity [23, 48, 49]. For water at 22 oC we have µ = 0.96 10−3pN s µm−2, which for L ≈ 10 µm results in ξt ≈ 0.0017 pN s µm−2 and ξn ≈ 0.0034 pN s µm−2. 21 References 1. Pazour GJ, Agrin N, Leszyk J, Witman GB. Proteomic analysis of a eukaryotic cilium. Science. 2005 Jul;170(1):103 -- 113. Available from: http://jcb.rupress. org/content/170/1/103.short. 2. Alberts et al B. Molecular Biology of the Cell. Garland Science, New York. 2002;. 3. Brokaw C. Direct measurements of sliding between outer doublet microtubules in swimming sperm flagella. Science. 1989 Mar;243(4898):1593 -- 1596. Available from: http://www.sciencemag.org/content/243/4898/1593.abstract. 4. Summers KE, Gibbons IR. Adenosine Triphosphate-Induced Sliding of Tubules in Trypsin-Treated Flagella of Sea-Urchin Sperm. Proceedings of the National Academy of Sciences. 1971 Jan;68(12):3092 -- 3096. Available from: http://www. pnas.org/content/68/12/3092. 5. Brokaw CJ. Molecular mechanism for oscillation in flagella and muscle. Proceedings of the National Academy of Sciences. 1975 Aug;72(8):3102 -- 3106. Available from: http://www.pnas.org/content/72/8/3102. 6. Julicher F, Prost J. Spontaneous Oscillations of Collective Molecular Motors. Physical Review Letters. 1997 Jun;78(23):4510 -- 4513. Available from: http: //link.aps.org/doi/10.1103/PhysRevLett.78.4510. 7. Camalet S, Julicher F. Generic aspects of axonemal beating. New Journal of Physics. 2000 Oct;2:24 -- 24. Available from: http://iopscience.iop.org/ 1367-2630/2/1/324. 8. Riedel-Kruse IH, Hilfinger A, Howard J, Julicher F. How molecular motors shape the flagellar beat. HFSP Journal. 2007;1(3):192 -- 208. Available from: http://www.tandfonline.com/doi/abs/10.2976/1.2773861. 9. Morita Y, Shingyoji C. Effects of imposed bending on microtubule sliding in sperm flagella. Current biology. 2004;14(23):2113 -- 2118. 10. Brokaw CJ. Bend Propagation by a Sliding Filament Model for Flagella. Journal of Experimental Biology. 1971 Jan;55(2):289 -- 304. Available from: http://jeb. biologists.org/content/55/2/289. 11. Brokaw CJ. Thinking about flagellar oscillation. Cell Motility and the Cytoskeleton. 2009 Aug;66(8):425 -- 436. Available from: http://onlinelibrary.wiley.com/ doi/10.1002/cm.20313/abstract. 12. Machin KE. Wave Propagation along Flagella. Journal of Experimental Biology. 1958 Dec;35(4):796 -- 806. Available from: http://jeb.biologists.org/content/ 35/4/796. 13. Lindemann CB. A model of flagellar and ciliary functioning which uses the forces transverse to the axoneme as the regulator of dynein activation. Cell Motility and the Cytoskeleton. 1994 Jan;29(2):141 -- 154. Available from: http: //onlinelibrary.wiley.com/doi/10.1002/cm.970290206/abstract. 14. Brokaw CJ. Thinking about flagellar oscillation. Cell motility and the cytoskeleton. 2009;66(8):425 -- 436. 22 15. Geyer VF, Sartori P, Friedrich BF, Julicher F, Howard J. The breaststroke beat of Chlamydomonas cilia is a spermlike flagellar wave that travels around a circular arc. In preparation. 2015;??(??):???? -- ???? 16. Satir P, Matsuoka T. Splitting the ciliary axoneme: Implications for a "Switch- Point" model of dynein arm activity in ciliary motion. Cell motility and the cytoskeleton. 1989;14(3):345 -- 358. 17. Everaers R, Bundschuh R, Kremer K. Fluctuations and stiffness of double-stranded polymers: railway-track model. EPL (Europhysics Letters). 1995;29(3):263. 18. Camalet S, Julicher F. Generic aspects of axonemal beating. New Journal of Physics. 2000;2(1):24. 19. Brokaw CJ. Bend propagation by a sliding filament model for flagella. Journal of Experimental Biology. 1971;55(2):289 -- 304. 20. Lauga E, Powers TR. The hydrodynamics of swimming microorganisms. Reports on Progress in Physics. 2009;72(9):096601. 21. Eshel D, Brokaw CJ. New evidence for a biased baseline mechanism for calcium- regulated asymmetry of flagellar bending. Cell motility and the cytoskeleton. 1987;7(2):160 -- 168. 22. Mukundan V, Sartori P, Geyer V, Julicher F, Howard J. Motor regulation results in distal forces that bend partially disintegrated Chlamydomonas axonemes into circular arcs. Biophysical journal. 2014;106(11):2434 -- 2442. 23. Johnson R, Brokaw C. Flagellar hydrodynamics. A comparison between resistive- force theory and slender-body theory. Biophysical journal. 1979;25(1):113. 24. Friedrich B, Riedel-Kruse I, Howard J, Julicher F. High-precision tracking of sperm swimming fine structure provides strong test of resistive force theory. The Journal of experimental biology. 2010;213(8):1226 -- 1234. 25. Camalet S, Julicher F, Prost J. Self-organized beating and swimming of internally driven filaments. Physical review letters. 1999;82(7):1590. 26. Alper J, Geyer V, Mukundan V, Howard J, et al. Reconstitution of flagellar sliding. Methods in enzymology. 2012;524:343 -- 369. 27. Ruhnow F, Zwicker D, Diez S. Tracking Single Particles and Elongated Fil- aments with Nanometer Precision. Biophysical Journal. 2011;100(11):2820 -- 2828. Available from: http://www.sciencedirect.com/science/article/pii/ S000634951100467X. 28. Machin K. Wave propagation along flagella. J Ex Biol. 1958;35(4898):796 -- 806. Available from: http://131.111.17.133/user/gold/pdfs/teaching/ machin1958.pdf. 29. Bayly P, Wilson K. Analysis of unstable modes distinguishes mathematical models of flagellar motion. Journal of The Royal Society Interface. 2015;12(106):20150124. 30. Sartori P, Geyer VF, Friedrich BF, Julicher F, Howard J. In a three-dimensional model of the axoneme transverse stresses can regulate the flagella beat in the presence of twist. In preparation. 2015;??(??):???? -- ???? 23 31. Brokaw CJ. Computer simulation of flagellar movement IX. Oscillation and symmetry breaking in a model for short flagella and nodal cilia. Cell motility and the cytoskeleton. 2005;60(1):35 -- 47. 32. Brokaw CJ. Computer simulation of flagellar movement VIII: Coordination of dynein by local curvature control can generate helical bending waves. Cell motility and the cytoskeleton. 2002;53(2):103 -- 124. 33. Brokaw C. Computer simulation of flagellar movement. VI. Simple curvature- controlled models are incompletely specified. Biophysical journal. 1985;48(4):633. 34. Bayly PV, Wilson KS. Equations of interdoublet separation during flagella motion reveal mechanisms of wave propagation and instability. Biophysical journal. 2014;107(7):1756 -- 1772. 35. Macnab RM, Koshland D. The gradient-sensing mechanism in bacterial chemotaxis. Proceedings of the National Academy of Sciences. 1972;69(9):2509 -- 2512. 36. Koshland Jr DE, Goldbeter A, Stock JB. Amplification and adaptation in regulatory and sensory systems. Science. 1982;217(4556):220 -- 225. 37. Yi TM, Huang Y, Simon MI, Doyle J. Robust perfect adaptation in bacte- rial chemotaxis through integral feedback control. Proceedings of the National Academy of Sciences. 2000;97(9):4649 -- 4653. 38. Shimizu TS, Tu Y, Berg HC. A modular gradient-sensing network for chemotaxis in Escherichia coli revealed by responses to time-varying stimuli. Molecular systems biology. 2010;6(1):382. 39. Witman G. The Chlamydomonas Sourcebook: Cell Motility and Behavior. vol. 3. Academic Press; 2009. 40. Smith EF, Yang P. The radial spokes and central apparatus: Mechano-chemical transducers that regulate flagellar motility. Cell motility and the cytoskeleton. 2004;57(1):8 -- 17. 41. Porter ME, Sale WS. The 9+ 2 axoneme anchors multiple inner arm dyneins and a network of kinases and phosphatases that control motility. The Journal of cell biology. 2000;151(5):F37 -- F42. 42. Lindemann CB. Structural-functional relationships of the dynein, spokes, and central-pair projections predicted from an analysis of the forces acting within a flagellum. Biophysical journal. 2003;84(6):4115 -- 4126. 43. Sartori P. Effect of curvature and normal forces on motor regulation of cilia. Technische Universitaet Dresden; 2015. 44. Cross MC, Hohenberg PC. Pattern formation outside of equilibrium. Reviews of modern physics. 1993;65(3):851. 45. Gittes F, Mickey B, Nettleton J, Howard J. Flexural Rigidity of Microtubules and Actin Filaments Measured from Thermal Fluctuations in Shape. The Journal of Cell Biology. 1993 Feb;120(4):923 -- 934. Available from: http://jcb.rupress. org/content/120/4/923. 46. Howard J. Mechanics of Motor Proteins and the Cytoskeleton. Sinauer Associates, Sunderland, MA. 2001;. 24 47. Nicastro D, Schwartz C, Pierson J, Gaudette R, Porter ME, McIntosh JR. The molecular architecture of axonemes revealed by cryoelectron tomography. Science. 2006;313(5789):944 -- 948. 48. Gray J, Hancock G. The propulsion of sea-urchin spermatozoa. Journal of Experimental Biology. 1955;32(4):802 -- 814. 49. Cox R. The motion of long slender bodies in a viscous fluid part 1. general theory. Journal of Fluid mechanics. 1970;44(04):791 -- 810. 25
1608.03730
1
1608
2016-08-12T09:12:28
A new look at blood shear-thinning
[ "physics.bio-ph", "cond-mat.soft" ]
Blood viscosity decreases with shear stress, a property essential for an efficient perfusion of the vascular tree. Shear-thinning is intimately related to the dynamics and mutual interactions of red blood cells (RBCs), the major constituents of blood. Our work explores RBCs dynamics under physiologically relevant conditions of flow strength, outer fluid viscosity and volume fraction. Our results contradict the current paradigm stating that RBCs should align and elongate in the flow direction thanks to their membrane circulation around their center of mass, reducing flow-lines disturbances. On the contrary, we observe both experimentally and with simulations, rich morphological transitions that relate to global blood rheology. For increasing shear stresses, RBCs successively tumble, roll, deform into rolling stomatocytes and finally adopt highly deformed and polylobed shapes even for semi-dilute volume fractions analogous to microcirculatory values. Our study suggests that any pathological change in plasma composition, RBCs cytosol viscosity or membrane mechanical properties will impact the onset of shape transitions and should play a central role in pathological blood rheology and flow behavior.
physics.bio-ph
physics
Classification: Physical Sciences - Applied Physical Sciences - Biological Sciences - Biophysics and Computational Biology A new look at blood shear-thinning Luca Lanotte1, Johannes Mauer2, Simon Mendez3, Dmitry A. Fedosov2, Jean-Marc Fromental4, Viviana Claveria1, Franck Nicoud3, Gerhard Gompper2, and Manouk Abkarian1 Authors affiliation: 1Centre de Biochimie Structurale, CNRS UMR 5048 - INSERM UMR 1054, University of Montpellier,34090 France 2Institute of Complex Systems and Institute for Advanced Simulation, Forschungszentrum Jülich, 52425 Germany 3IMAG UMR 5149 CC 051, University of Montpellier, 34095 Montpellier 4Laboratoire Charles Coulomb, CNRS UMR 5587, University of Montpellier,34095 France Corresponding author: Manouk ABKARIAN Email: [email protected] 6 1 0 2 g u A 2 1 ] h p - o i b . s c i s y h p [ 1 v 0 3 7 3 0 . 8 0 6 1 : v i X r a 1 A new look at blood shear-thinning August 15, 2016 Blood viscosity decreases with shear stress, a property essential for an effi- cient perfusion of the vascular tree. Shear-thinning is intimately related to the dynamics and mutual interactions of red blood cells (RBCs), the major con- stituents of blood. Our work explores RBCs dynamics under physiologically relevant conditions of flow strength, outer fluid viscosity and volume fraction. Our results contradict the current paradigm stating that RBCs should align and elongate in the flow direction thanks to their membrane circulation around their center of mass, reducing flow-lines disturbances. On the contrary, we observe both experimentally and with simulations, rich morphological transi- tions that relate to global blood rheology. For increasing shear stresses, RBCs successively tumble, roll, deform into rolling stomatocytes and finally adopt highly deformed and polylobed shapes even for semi-dilute volume fractions analogous to microcirculatory values. Our study suggests that any patho- logical change in plasma composition, RBCs cytosol viscosity or membrane mechanical properties will impact the onset of shape transitions and should play a central role in pathological blood rheology and flow behavior. 2 Significance Statement Blood viscosity decreases with shear stress, a property essential for an effi- cient perfusion of the human body. This shear-thinning property is intimately related to the dynamics of its major constituents, the red blood cells (RBCs). Our work explores this dynamics under physiologically relevant conditions of both flow and viscosity and contradicts the current paradigm describing RBCs behavior as the one of viscous droplets. Using both experiments and simulations, our study shows that the lack of membrane fluidity is the key feature which controls dynamic morphological transitions and global rheol- ogy. Our study questions the physiological relevance of previous droplet-like descriptions of RBCs dynamics. 1 Introduction Red blood cells (RBCs) are the main cellular component of whole blood (WB). About thousand times more concentrated than white blood cells and platelets, the volume frac- tion of RBCs, also called the hematocrit (Ht), is as high as 45%. Local dynamics and interactions of RBCs govern blood rheology and control proper perfusion of the entire vascular tree. In particular, WB exhibits a strong shear-thinning behavior in shear flow [1, 2] which is determined by the aggregability and deformability of RBCs [3, 4], since the suspending plasma behaves essentially as a Newtonian fluid with a constant shear viscosity of about 1.2 cP at 37◦C [1, 5]. Macromolecules dispersed in the plasma, such as fibrinogen, induce an attractive force between RBCs leading to their aggregation [6, 7]. Because RBCs have a biconcave disk-like shape at rest, they form long floppy rouleaux structures which can reversibly and continuously break down to single flowing discocytes 3 for increasing shear rates γ up to tens of s−1 [8, 9]. This change in microstructure is indeed accompanied by a strong viscosity drop from ∼ 100 cP down to about ∼ 10 cP [10]. However, for shear rates between 10 and 1500 s−1, which are very common in the microcirculation [11, 12], only the deformability and dynamics of RBCs can account for a further 4-fold decrease in WB viscosity down to values as low as 2 − 3 cP [3, 4]. This value strongly depends on the hematocrit and the viscosity of the hemoglobin cytoplasm of RBCs, which is around 5 cP at 37◦C [13]. The link between cellular deformability and shear-thinning remains unsettled, even though it is crucial in understanding blood flow both in health and disease, since altered deformability of RBCs [14] has indeed been correlated with impaired perfusion, increase in blood viscosity, and vaso-occlusion in microcirculatory disorders such as diabetus mel- litus, or hemoglobinopathies like sickle cell anemia. Measurements of WB rheology in microtubes [4, 15] and rheoscopes [16, 17], realized during the 1970s and the 1980s, have led to the foundation of the current paradigm for shear-thinning, comforting an emulsion analogy inherited from the 1960s [1, 2]. While at small shear stresses (less than about 5× 10−2 Pa), single RBCs flip like a coin, they attain a steady orientation with respect to the direction of the flow for increasing shear stresses [16, 17]. This concept is correct only when RBCs are dispersed into a viscous aqueous solution (often of dextran polysaccha- rides), which is several times more viscous than the inner cytoplasm, as has been done in most of available experiments [16, 17]. Such a viscous environment is supposed to mimic the effective high-shear-rate viscosity of the WB suspension measured by rheometry [18]. Single RBCs in such viscous solutions behave similar to liquid droplets. Their composite membrane elements rotate around their center of mass as the tread of a tank, transfer- ring to the cytoplasm the vorticity necessary for orientation stabilization [17]. Membrane tank-treading with a steady alignment would effectively assist the flow by minimizing the 4 disturbances to flow lines. Subsequent gradual elongation of the discocytes into prolate ellipsoids for shear stresses increasing up to few Pa would then explain the deformation- and shape-controlled shear-thinning behavior of WB [9]. Therefore, an essential element of this paradigm is that single RBCs should tank-tread in response to high shear stresses. Recent experiments on dilute RBC suspensions, realized with outer viscosities similar to that of the hemoglobin-rich cytoplasm, have demonstrated however that the solid-like tumbling motion of isolated RBCs is not replaced by the tank-treading dynamics for increasing shear rates, but rather by a typical rolling motion, where the axis of symmetry of the discocyte lies in the direction of the vorticity [19]. Even for shear stresses as high as 0.1− 0.5 Pa, no fluidization of the membrane was observed, demonstrating the important role the inner-to-outer viscosity ratio λ plays in local dynamics. In fact, the viscosity of blood plasma under physiological conditions is about five times smaller than that of the RBC cytosol. Hence, new studies are required to explore the local dynamics of RBCs at such low viscosity conditions, especially when γ is greater than 10 s−1 or in the range of stresses between a tenth to a few Pa, which are characteristic for microcirculatory flow [11, 12]. In this work, we explore the shape and dynamics of RBCs in the range of shear rates and hematocrits relevant for the microcirculation and correlate local RBCs morphologies with global shear-thinning behavior measured by rheometry. Using microfluidics, we demonstrate that RBCs show unexpected dynamics characterized by rotating polylobed shapes, with no clear tank-treading of their membrane. Experimental results obtained by both microfluidic observations and rheological measurements are corroborated by two different 3D simulation techniques: a continuum approach YALES2BIO [20] based on the finite-volume method and a mesoscopic approach for blood-flow modeling [21] based on the smoothed dissipative particle dynamics (SDPD) method [22] (see Materials and 5 Methods). We discuss the implication of this work for microcirculatory flow and re- interpret the classical shear-thinning conjecture. 2 Results 2.1 Morphology of RBCs in simple shear flow We explore first the morphologies of single RBCs at 37◦C submitted to a simple shear flow in a range of γ between 10 s−1 and 2000 s−1, using a cone-and-plate configuration of a rheometer. Cells are hardened with an aldehyde treatment while flowing (see protocol in Materials and Methods). Samples of these hardened RBCs (HRBCs) are then with- drawn from the plate of the rheometer and visualized by bright field (BF) and confocal microscopy after fluorescent dying of their membranes. Typical images of the fixed shapes are shown in fig. 1A for a given sample. We classify the different morphologies of HRBCs as a function of γ in fig. 1B using an in-house image-analysis program. RBCs subjected to weak shear rates (< 10 s−1) commonly keep their biconcavity sim- ilar to typical discocytes at rest (fig. 1A). For γ increasing from 10 to 40 s−1 however, the percentage of discocytes decreases more than twice (fig. 1B). This is associated with the systematic development of two populations of cup-shaped stomatocytes: one with a circular rim and another with an elliptical one with an aspect ratio smaller than 0.95. Examples of these deformed stomatocytes are displayed in fig. 1A at 18 and 40 s−1, respectively. At γ = 45 s−1, the fraction of discocytes has dramatically dropped repre- senting only about 30% of the total population, while the total amount of stomatocytes has jumped nearly to 65% of the sample. These values remain nearly constant for both cell populations up to about 400 s−1 (light blue color region of fig. 1A). However, in the region between 45 s−1 and 400 s−1, the fraction of circular-rim-shaped stomatocytes decreases, and gets gradually replaced by the population of elliptical-rim-shaped stomatocytes with 6 an increased ellipticity. At γ ≈ 400 s−1, a new substantial transition takes place. As the shear rate goes beyond 400 s−1, we observe a sharp increase in the population of deformed stomatocytes which appear to be strongly folded (image for 475 s−1 in fig. 1A). Most importantly, we also see the rise of a new population of RBCs showing large lobes on their surface. In most cases, we find cells with three or six lobes forming tetrahedra, which will be henceforth referred to as trilobes or hexalobes, respectively. In the region 400 s−1 < γ < 2000 s−1, marked by a light red color in fig. 1B, we observe the nearly disappearance of discocytes (less than 10%) and the formation of trilobes and to a lesser degree hexalobes. At γ = 2000 s−1, the polylobed RBCs represent almost 70% of the sheared population. Moreover, rapid rise of hexalobe population is found for γ ∼ 2000 s−1, suggesting a direct correlation between the number of lobes and shear strength. 2.2 RBCs dynamics in flow To understand further how the acquired morphologies couple to the flow and to rule out hardening artifacts, we perform complementary microfluidic experiments with high-speed video microscopy at 25◦C using dilute suspensions of RBCs (Ht ≈ 1%) in PBS/BSA only (no hardening). The suspension is fed at increasing flow rates into a circular cross- section microcapillary for the same range of shear rates explored previously by rheometry. Local shear rate is evaluated by measuring both the local cell velocity and its distance from the capillary walls. Different dynamics are detected for increasing values of this estimated γ. Their time sequences are shown in fig. 2A obtained in a field of view of about 300 µm long. At low shear rates ( γ < 40 s−1), the tumbling-to-rolling transition is observed, where the axis of symmetry of discocytes gradually aligns with the vorticity direction. As in the rheological setup, the number of rolling RBCs becomes significant for 7 increasing shear rates. In addition, a substantial increase in the population of cup-shaped stomatocytes is clearly detected, which show a rolling motion in the flow as illustrated in fig. 2A at γ = 150 s−1. Moreover, observations reveal both tumbling and vacillating- breathing behaviors (see Movie S1) of folded stomatocytes (fig. 2A, blue box) with the latter being largely detected at high shear strengths γ > 350 s−1 (see examples in fig. 2 for 250 s−1and 500 s−1). Furthermore, these in vitro investigations confirm the formation and stability of polylobed shapes at higher shear rates. Two trilobes are depicted in side and top views at 750 and 800 s−1, respectively, in fig. 2A (light red color box) and in Movie S2. At such high shear rates, RBCs elongate in the vorticity direction, and display three lobes which rotate around the center of mass. An even more solid-like rotation is observed for hexalobes under comparable flow conditions (last time sequence in fig. 2A and in Movie S3). These experiments show that no sharp morphological transition occurs for increasing shear rates, but rather a marked variation in shape distribution of RBC populations is present. Finally, thanks to our pressure injection system, we can produce very fast "stop flow" experiments. These experiments nicely demonstrate the dynamical nature of polylobed morphologies, since either trilobes or hexalobes return rapidly to their resting discocyte shape, when the flow is stopped, as depicted in fig. 2B and in Movie S4. Noteworthy, the population of stomatocytes keeps its shape longer and relaxes to the discocyte population on a longer time scale of tens of minutes. To complement the experiments, we also perform numerical simulations of isolated RBCs in simple shear flow for varying shear rates, using both SDPD and YALES2BIO softwares (see Materials and Methods). The simulated dynamics is presented in fig. 2A using only YALES2BIO results, since agreement between the two simulations is very good (some supplementary movies in SDPD are also provided). Simulations confirm the sequence of shapes and dynamics observed experimentally as well as their dependence 8 on γ as shown in fig. 2A. In comparison to experiments, simulations provide a detailed information about dynamic behavior of RBCs in flow. For example, the same cell may have different stable shapes under flow (observed during several hundreds of time units 1/ γ) depending on its initial orientation with respect to the flow. In particular, stomato- cytes can display both rolling-like or tumbling-like motion; rolling deformed stomatocytes (see Movie S5) and tumbling trilobes (see Movie S6) might be found for the same shear rate. However, the sequence of shapes found experimentally appears to be qualitatively robust to moderate changes in the RBC mechanical properties, which has been tested in simulations. One difference between experimental and simulation results is hexalobe deformation of RBCs (fig. 2A at the bottom), which has been observed in simulations only as a transient state. Nevertheless, in none of the simulations performed for the range 1 s−1< γ < 2000 s−1 a tank-treading ellipsoidal RBC has been obtained. 2.3 Morphology and dynamics of RBCs at high hematocrits To evaluate the prevalence of polylobed shapes at physiologically relevant shear rates and hematocrits, same hardening experiments done in flow and described earlier were realized for several RBC suspensions with different Ht values of 5, 15, 22, 35 and 45%. γ = 900 s−1 (at 37◦C) has been selected for these experiments, since the analysis for dilute sus- pensions in fig. 1B has shown a maximum in the population of trilobes at this shear rate. Figure 3A indicates a significant decrease in the probability to obtain polylobed HRBCs for increasing Ht. The percentage of trilobes and hexalobes, which amounts for about 55% of the entire population at Ht = 5%, is more than halved when the hematocrit reaches 45%. In addition, at high enough Ht many HRBCs display a deformed shape with multiple irregular lobes, making it difficult to precisely classify their morphology; these cells will be called "multilobes" with an illustration in the lower inset of fig. 3A. The 9 fraction of these "multilobed" RBCs seems to be rather stable amounting for 25% − 35% of the total population. However, the onset of a new cell morphology is detected for increasing Ht. Flattened discocytes characterized by one or several grooves or creases on their membrane have been found and are illustrated in the inset of fig. 3A by optical and confocal microscopy. The fraction of creased discocytes increases significantly at high Ht values amounting for almost 50% of the sample at Ht = 45%. In contrast to the vast majority of polylobed RBCs in dilute suspensions, the formation of flat cells with creases is a clear effect of mutual hydrodynamic interactions between cells in such dense suspensions. Indeed, the total percentage of lobed cells with increasing Ht has the exact inverse evolution of that for creased discocytes as shown in fig.3A. Interestingly, physio- logical hematocrits in the microcirculation lie in the range 15 − 30% (light yellow region in fig.3A), in which a clear coexistence of different cell morphologies with a domination of lobed shapes is present. To corroborate the existence of such folded and irregular structures in flow at high Ht, we have performed SDPD simulations with dense RBCs suspensions. The analysis of creased discocytes population for different Ht is presented in fig. 3A along with the cor- responding experimental data showing a good qualitative agreement. Figure 3B presents RBC-extension distributions for different Ht values, which are averaged over multiple RBCs and many time instances (A typical simulation is shown in Movie S7). The RBC extension corresponds to the maximum instantaneous cell length in the flow direction, which is illustrated in one of the insets in fig. 3B. The extension distributions clearly show that RBCs become more stretched in the flow as Ht is increased. The two regions indicated by light yellow and light blue colors in fig. 3B schematically illustrate the ranges of cell extensions with polylobed shapes and creased discocytes. Thus, at low enough Ht polylobed shapes dominate, while at Ht = 45% the fraction of creased discocytes becomes 10 larger than one half. The coexistence of various cell morphologies and the evolution of different RBCs populations with increasing Ht are fully consistent with the observations made experimentally. 2.4 The role of cell deformability in shear-thinning rheograms Now, we re-examine the interpretation of classical shear-thinning rheograms for human blood, which were introduced in the 1970s [3] and provided the seminal link between shear- induced RBC deformation, tank-treading, and shear-thinning. We perform experiments at Ht = 45% suspending RBCs in either PBS/BSA or their native plasma. The samples are sheared in the rheometer at 37◦C (using both Couette and cone-and-plate geometry) for the range of γ between 5 s−1 and 2000 s−1. The relative viscosity (ηrel),defined as the ratio between the measured viscosity of the suspension (ηs) and the viscosity of the suspending medium (ηm), is obtained for different γ. Consistently with previous works [3], at low shear rates, washed blood has a significantly lower viscosity than WB since plasma induces aggregation. From γ ≈ 10 s−1, viscosity curves for both washed blood and WB merge into a common shear-thinning line as shown in fig. 4A. To understand the role of shape change in this regime, we compare washed blood rheology with that of two HRBC suspensions at 45% Ht: one with stiff discocytes hardened at rest and the other one with polylobed cells obtained by hardening cells previously in the rheometer at γ = 1500 s−1 and resuspending them at the desired 45% Ht. Both suspensions of HRBCs show a nearly Newtonian behavior except for shear rates above 400 s−1 where the onset of a slight shear-thickening is observed especially for hardened discocytes. This behavior is similar to shear-thickening reported for rigid colloidal suspensions [23, 24]. For both HRBC dispersions viscosity values are larger than those for washed blood; however, the sample with hardened polylobes yields a 70% lower relative viscosity than that for 11 stiffened discocytes, as seen in Fig 4A. These results indicate that from γ = 10 s−1 up to a few hundred s−1, shear-thinning is largely determined by the change in RBC dynamics from tumbling to rolling and by the deformation of RBCs into elongated stomatocytes. Beyond γ ∼ 400 s−1 however, a further viscosity drop is related to the formation of polylobed and flattened shapes discussed above. SDPD simulations performed for λ = 5 and Ht = 45% quantitatively capture shear- thinning as shown in fig. 4A. A closer look at RBCs shapes for different shear rates essentially confirms experimental observations. For moderate shear rates up to a few hundred s−1, a prevalence of stomatocyes and multilobe cells is found, while at high shear rates, flattened RBCs are mainly observed. Stomatocytes and polylobes found for moderate γ values show a tumbling-like dynamics. The dynamics of flattened RBCs at high shear rates resembles a tank-treading motion; however, these cells display persistent and dynamic membrane deformations represented by small creases or larger lobes, which continuously form and disappear (see Movie S7). Tank-treading with a steady ellipsoidal shape was never observed in simulations and the motion of flattened RBCs at high shear rates is likely to be a superposition of both cell stretching and lobes rotation. Finally, to highlight the importance of λ for the presence of lobular shapes and their role in shear-thinning in comparison to steady tank-treading, we perform rheological mea- surements for suspensions of normal RBCs dispersed in solutions with increasing viscosi- ties modulated by various dextran concentrations of 2%, 4% and 9% (wt/wt). Hematocrit is kept at 45% and the suspensions are sheared at 25◦C for the range of γ between 10 s−1 and 2000 s−1 (without hardening). Figure 4B presents relative viscosities of these suspensions as a function of the shear stress τ (= ηs γ). The curves for PBS/BSA and 9% dextran solution display inconsistent trends. In PBS/BSA, which corresponds to high viscosity ratios (λ ≈ 5 − 6), a significant and continuous decrease of ηrel is observed for 12 increasing τ up to the rheometer limit of about 6 Pa. In contrast, 9% dextran suspen- sions (λ ≈ 0.15) exhibit a slight shear-thinning for stresses close to 1 Pa, but ηrel yields a plateau for higher stresses (see fig. 4B). Samples with 2% and 4% dextran solutions show an intermediate behavior, in which the occurrence of an increasingly significant shear-thinning for increasing λ is detected. In parallel, we have performed microfluidic observations of single RBCs flowing in 2% and 9% dextran solutions at Ht = 1% for comparable shear stresses. Contrary to the behavior of RBCs in PBS/BSA, cells in 9% dextran solutions do not assume polylobed shapes, but show a rather sharp transition from rigid tumbling discocytes to tank-treading ellipsoids which elongate with increasing stresses. This transition takes place at τ ∼ 1P a which is consistent with the change in slope of the corresponding ηrel curve in fig. 4B. In the inset of fig. 4B an illustration of the effect of λ on RBC dynamics is presented. Images show the shapes of cells dispersed respectively in PBS/BSA, 2% and 9% dextran solutions and flowing in a microchannel with the same estimated local shear stress of τ = 6 Pa. A decrease in λ promotes in-flow orientation, elongation and simultaneous reduction of lobes size (see Movie S8). When λ ≈ 0.15, polylobed cells are completely replaced by elongated tank-treading ellipsoids (see Inset fig. 4B). Noteworthy, all images in the inset of fig. 4B and movies in the Supplemental Materials refer to RBCs flowing in microchannels in which the presence of a gradient in γ can potentially affect RBCs dynamics, while rheological measurements were performed under pure shear flow conditions. However, comparison between off-center shape sequences in 50 µm capillaries and simple-shear YALES2BIO simulations presented in fig. 2A indicates a rather marginal effect of the gradient of shear rate.. 13 3 Discussion Investigation of blood rheology finds its cornerstone in the seminal work of Chien and collaborators [3]. These authors have shown how the deformability of RBCs influences blood viscosity, especially at high shear rates and hematocrits. "Deformability" means that cells can change their shape in response to mechanical stresses. This term encom- passes both out-of-plane and in-plane deformations of membrane elements. In the 1970s, deformability has been mainly attributed to tank-treading observed when RBCs were sus- pended in a viscous fluid mimicking the viscosity of WB at high shear rates. The resulting stabilization of cells' orientation implied a gradual elongation of the cells for increasing shear rates, which was believed to be the main reason for shear-thinning. This droplet-like behavior of RBCs led to the analogy between blood and emulsion rheology [16, 17]. Our morphological and rheological studies in combination with microfluidic observa- tions and simulations lead to a different picture. Without aggregation, blood continuous shear-thinning is the result of a rich dynamical behavior of RBCs population, which is based primarily on the inability of RBCs to tank-tread under physiological conditions (λ ≈ 5). We have observed a diversity of shapes and dynamical states of RBCs for increasing shear rates and the development of a large fraction of highly deformed and polylobed cells for γ > 400 s−1 (figs. 1 and 2) which has been confirmed by simulations. We first discuss single-cell behavior in dilute suspensions. For γ < 1 s−1, the biconcave shape is preserved and the behavior of RBCs is similar to that of a rigid oblate ellipsoid, such that the membrane and the enclosed fluid rotate as a rigid body. For increasing shear rates up to 10 s−1, more and more cells are found to roll on their edge as previ- ously observed in experiments using rheoscopes [25] and more recently in flow chambers [26, 19]. Shear-thinning in this range of shear rates is therefore mainly controlled by 14 discocyte orientation. Rolling has its origin in shear elasticity of the membrane. Under physiological conditions, membrane tank-treading is forbidden due to the relatively high internal viscosity as predicted by viscous ellipsoidal models of the RBC [27]. However, during each tumbling period the membrane elements oscillate around a given position and this local oscillatory strain seems to destabilize RBCs from tumbling toward rolling [28] instead of tank-treading. Though not fully settled, this phenomenon is well captured by our simulations and is described as a stable motion in several recent numerical simulations of capsules [29] and RBCs [30]. In the range of shear rates between 45 and 400 s−1, a large population of the rolling discocytes looses one dimple and becomes rolling stomatocytes displaying at first more frequently a circular rim up to 45 s−1 and then a breathing elliptical rim for shear rates reaching 400 s−1. In this second morphological transition, rolling stomatocytes are show- ing a smaller cross-section to shear flow than rolling discocytes, reducing their distur- bances of the flow lines. This leads to a further shear-thinning. Both our numerical approaches capture this flow-induced morphological transition for single RBCs. The loss of the dimple happens abruptly in simulations highlighting a buckling instability, which is also confirmed experimentally by the long time scales (minutes) necessary for the pop- ulation of stomatocytes to relax to a discocyte shape. Beyond 400 s−1, the dilute population of RBCs is subjected to a new shift in mor- phology and large lobes develop on the surface of the cells (fig. 1A). These out-of-plane deformations are again a signature for a lack of fluidization of the membrane even for the largest stress of 6 Pa we were able to reach both experimentally (with the rheometer and in microflows) and numerically. Noteworthy, Sutera and Mehrjardi [31] studied in the 1970s the morphology of cells hardened in PBS solutions at very high shear stresses (τ ≥ 10 Pa). Their images at τ = 10 Pa show the presence of pronounced concavities on RBCs 15 similar to those we observe for polylobed cells. Thus, trilobes or even hexalobes are the result of the large viscosity contrast. Our experiments with varying viscous environment confirm how lobes are directly associated with a large enough viscosity contrast λ since their presence vanishes in very viscous dextran solutions as shown in fig. 4. However, the exact nature of these transitions and the development of folded shapes remain to be understood. Both experiments and simulations suggest that large enough viscosity ratio is at least sufficient for polylobes to appear. However, modifications of a RBC membrane at rest can also induce similar transformations observed here for increasing shear rates. Theoretically, a large equilibrium-shape variety has been obtained for RBCs including stomatocytes, with circular and deformed rims, and trilobes (named knizocytes in the lit- erature), by changing the resting elastic state of the cytoskeleton and the area difference between lipid leaflets in their model [32]. Despite this striking similarity, the dynamic shapes in simulations under flow have been observed without any explicit modification of the membrane. These results open the road for further experimental investigations to explore the effect of flow on the shape modification of a RBC membrane. Our dilute single-cell experiments and simulations show how out-of-plane deformations are inevitable for increasing shear rates. However, for more concentrated suspensions and as hematocrit is increased, a large population of creased and flattened discocytes appears both in experiments (see Movie S9) and simulations (see for instance Movie S7 in supplementary information and fig. 3B). While it is perfectly conceivable that large lobes cannot easily develop on membranes in crowded suspensions, the complex flow between cells seems to counter-balance the inherent tendency of RBCs to elongate more in the vorticity direction. It is this fine balance between vorticity- and flow-direction elongations that aligns cells on average parallel to the flow. Nevertheless, any variations in relative positions between the cells or any collisions immediately give birth to local bulges 16 on the membranes and produce large fluctuations of shapes as seen in RBCs extension distributions of the simulation presented in fig. 3B. In conclusion, our study reveals that blood shear-thinning is related to a rich behavior of RBCs in shear flow convoluted with a large distribution of cell shapes for any given flow condition. The lack of membrane fluidity for high viscosity contrast between inner and outer fluids is the key feature which controls RBCs behavior. As a consequence, the droplet-based analogy for blood rheology at high shear rates appears to be erroneous for the explanation of shear-thinning. Moreover, several fundamental physiological phe- nomena have been analyzed under the assumption of membrane tank-treading, such as vaso-regulatory ATP release by RBCs in strong shear flows [33] or the formation of a few-micron thick cell-free layer adjacent to the vessels walls which is responsible for the apparent viscosity drop with decreasing vessel diameter (the so-called Fåhraeus-Lindqvist effect [34]). Our study questions the relevance of a droplet-like analogy for RBC dynamics to explain these phenomena and asks to re-examine them both experimentally and the- oretically for physiologically relevant viscosity and stress conditions. Finally, our study suggests that in pathological change in plasma composition, RBCs cytosol viscosity or membrane mechanical properties, will impact the onset of shape transitions and should play a central role in pathological blood rheology and flow behavior. 4 Materials and Methods 4.1 Sample preparation Fresh venous-blood samples were obtained from healthy consenting donors, thanks to the agreement with a local blood bank to use blood (EFS) and its components for non- therapeutic purposes. The samples, whose volume ranged between 420 and 500 ml, were stored in PVC-DEHP transfusion bags produced by Macopharma (MSE Systems). Such 17 bags are provided with 66.5 ml of CPD anticoagulant, whose composition is: Sodium citrate dihydrate 26.3 g/l, Citric acid monohydrate 3.27 g /l, Monobasic Sodium Phos- phate dihydrate 2.51 g/l, Dextrose monohydrate 25.5 g/l. The suspensions were then centrifuged at 1000 x g for 20 min and the buffy coat was removed. RBC dispersions were diluted in PBS and the procedure of centrifugation/removal was repeated twice to obtain washed RBCs. Afterwards, cells were re-suspended in their native plasma treated with anticoagulant or in solutions of PBS supplemented with BSA (1 mg of BSA per 1 ml PBS) and the hematocrit was adjusted by using graduated and narrow test-tubes. Samples were kept at 25◦C throughout all the preparation. Dilute suspensions (Ht = 1%) of fluorescent HRBCs for confocal observation were prepared by using a PKH26 Red Fluorescent Cell Linker (Sigma-Aldrich). HRBCs were labeled following standard protocol of the provider. 2,000 kDa Dextran (Leuconostoc spp., Sigma-Aldrich) solutions were prepared in PBS with a concentration of 2% (wt/wt), 4% (wt/wt), and 9% (wt/wt). Dissolution was obtained at room temperature by overnight gentle stirring. All dextran/PBS solutions showed a nearly Newtonian behavior with a constant viscosity equal to 3.4, 10.2 and 38.8 mPa*s respectively for dextran concentrations of 2% (wt/wt), 4% (wt/wt), and 9% (wt/wt). We used a very high molecular weight dextran in order not to change the osmolarity of the buffer suspension when it has been added to make the suspending medium more viscous. All dispersions containing RBCs and all suspending solutions used in our experiments have shown a pH ranging between 7.2 and 7.5 and osmolarities of our buffers were measured to be 285 mOsm on average. The samples were stored in a refrigerator at 4◦C and they were used within the time frame between 3 hours and 5 days after the blood draw. The healthy state of the cells have been checked by looking at the amount of echinocytes before each experiments. (Blood samples and realized experiments 18 are summarized in table 1 of the supplementary text). 4.2 Rheology and microfluidics Rheology experiments were performed in a cone-plate stress-imposed rheometer (AR 2000 - TA Instrument). The temperature was kept constant at 37◦C by a Peltier temperature controller. Data were analyzed by the Rheology Advantage Data Analysis software. In the first type of experiments, diluted suspensions of RBCs (Ht=1%) in a PBS/BSA solution were hardened in flow by adding a 70% wt/wt solution of glutaraldehyde prepared in PBS solution (Sigma-Aldrich) and glutaraldehyde concentration was adjusted to 9%. 300 µl of the resulting sample was added by capillarity to 700 µl of the RBC dispersion during shearing to reach a final glutaraldehyde concentration of almost 3%. Such an amount of glutaraldehyde produces a very rapid, strong and permanent solidification. Indeed, recent AFM indentation measurements [35] performed on glutaraldehyde-treated RBCs have yielded a Young's modulus E in the range 10-100 kPa. Therefore, for the maximum shear stress σ of 6 Pa reached in our experiments, the typical strain of hardened RBCs, σ/E, is expected to be within 0.05-0.005%. This means that HRBCs can be considered as rigid objects in all our experiments. The same protocol was used to harden samples at higher hematocrits. Both shearing and hardening were carried out within 2 min. Samples were then collected, washed, and suspended in PBS solution for microscopic examination. Images of stiffened cells were acquired with an inverted microscope (Olympus IX71 60x objective) equipped with a digital camera (Sony XCD-X710). Acquired data were analyzed in Matlab (Mathworks, version 2011a): the minimum number of cells counted and classified for each γ in fig. 1B and for each Ht in fig. 3A was equal to 550 and 450, respectively. Microflow observations were realized at 25◦C in glass round capillaries with a diameter of 50 µm. Flow rates were adjusted by a OB1 Pressure Controller (ElveFlow) 19 operating up to 2 bar. Images were acquired at a high magnification (60x and 100x objectives) on a Nikon inverted microscope equipped with a high-speed camera Phantom Miro M320S (Vision Research) and subsequently analyzed by a custom image analysis software (ImageJ). 4.3 Numerical simulations Two different in-house softwares were used for numerical simulations. The first one, YALES2BIO [20, 36], is a finite-volume parallel solver for the incompressible Navier-Stokes equations on unstructured meshes. Fluid-structure coupling was implemented using an immersed boundary method adapted to unstructured grids [20]. RBCs were modeled as viscous drops enclosed by membranes resisting shear, bending, and area dilation. In the SDPD simulations, the smoothed dissipative particle dynamics method [22], a mesoscale hydrodynamic particle-based approach, represents fluid flow, while a RBC membrane was modeled as a triangulated network of springs [21], whose vertices are coupled to the fluid via frictional forces. The network assumes fixed connectivity and includes the spring's elastic energy, bending energy, and area and volume conservation constraints [21]. Further details about the numerical methods and setups can be found in supplementary information. Acknowledgments M.A., S.M., F.N. and L.L. acknowledge the support of the OSEO-BPI project Dat@Diag. M.A. and L.L. are grateful for the support from the Labex Numev (convention No. ANR- 10-LABX-20). S.M. and F.N. acknowledge the support of the ANR (ANR-11-JS09-0011) and thank Dr. Moureau and Dr. Lartigue (CORIA, UMR 6614) and the SUCCESS scien- tific group for providing the YALES2 code, which served as a basis for the development of 20 YALES2BIO. Simulations with YALES2BIO were performed using HPC resources from GENCI-CINES (Grant 2015-c2015037194). D.A.F. acknowledges funding by the Alexan- der von Humboldt Foundation. SDPD simulations were performed using a CPU time allocation at the Jülich Supercomputing Center. References [1] Wells, RE, Merrill, EW (1961) Shear rate dependence of the viscosity of whole blood and plasma. Science 133:763–764. [2] Dintenfass, L (1968) Internal viscosity of the red cell and a blood viscosity equation. Nature 219:956–958. [3] Chien, S (1970) Shear dependence of effective cell volume as a determinant of blood viscosity. Science 168:977–979. [4] Skalak, R, Keller, SR, Secomb, TW (1981) Mechanics of blood flow. J. Biomech. Eng. 103:102–115. [5] Brust, M et al. (2013) Rheology of human blood plasma: viscoelastic versus Newto- nian behavior. Phys. Rev. Lett. 110:078305. [6] Merrill, EW, Gilliland, ER, Lee, TS, Salzman, EW (1966) Blood rheology: effect of fibrinogen deduced by addition. Circ. Res. 18:437–446. [7] Wagner, C, Steffen, P, Svetina, S (2013) Aggregation of red blood cells: from rouleaux to clot formation. C. R. Physique 14:459–469. 21 [8] Qin, Z, Durand, LG, Allard, L, Cloutier, G (1998) Effects of a sudden flow reduction on red blood cell rouleau formation and orientation using RF backscattered power. Ultrasound Med. Biol. 24:503–511. [9] Fedosov, DA, Pan, W, Caswell, B, Gompper, G, Karniadakis, GE (2011) Predicting human blood viscosity in silico. Proc. Natl. Acad. Sci. USA 108:11772–11777. [10] Schmid-Schönbein, H, Wells, RE (1971) Rheological properties of human erythro- cytes and their influence upon the "anomalous" viscosity of blood. Ergeb. Physiol. 63:146–219. [11] Lipowsky, HH (2005) Microvascular rheology and hemodynamics. Microcirculation 12:5–15. [12] Popel, AS, Johnson, PC (2005) Microcirculation and hemorheology. Annu. Rev. Fluid. Mech. 37:43–69. [13] Wells, R, Schmid-Schönbein, H (1969) Red cell deformation and fluidity of concen- trated cell suspensions. J. Appl. Physiol. 27:213–217. [14] Diez-Silva, M, Dao, M, Han, J, Lim, CT, Suresh, S (2010) Shape and biomechanical characteristics of human red blood cells in health and disease. MRS Bulletin 35:382– 388. [15] Goldsmith, HL, Marlow, J (1972) Flow behaviour of erythrocytes. I. Rotation and deformation in dilute suspensions. Proc. R. Soc. Lond. B 182:351–384. [16] Schmid-Schoenbein, H, Wells, R (1969) Fluid drop-like transition of erythrocytes under shear. Science 165:288–291. 22 [17] Fischer, TM, Stöhr-Lissen, M, Schmid-Schönbein, H (1978) The red cell as a fluid droplet: tank tread-like motion of the human erythrocyte membrane in shear flow. Science 202:894–896. [18] Fischer, TM (1977) Tank tread motion of red-cell membranes in viscometric flow- behavior of intracellular and extracellular markers (with film) (Blood Cells) Vol. 3. [19] Dupire, J, Socol, M, Viallat, A (2012) Full dynamics of a red blood cell in shear flow. Proc. Natl. Acad. Sci. USA 109:20808–20813. [20] Mendez, S, Gibaud, E, Nicoud, F (2014) An unstructured solver for simulations of deformable particles in flows at arbitrary Reynolds numbers. J. Comp. Phys. 256:465–483. [21] Fedosov, DA, Caswell, B, Karniadakis, GE (2010) A multiscale red blood cell model with accurate mechanics, rheology, and dynamics. Biophys. J. 98:2215–2225. [22] Müller, K, Fedosov, DA, Gompper, G (2015) Smoothed dissipative particle dynamics with angular momentum conservation. J. Comp. Phys. 281:301–315. [23] Barnes, HA (1989) Shear-thickening (dilatancy) in suspensions of non aggregating solid particles dispersed in newtonian liquids. Journal of Rheology 33:329–366. [24] Brown, E et al. (2011) Shear thickening and jamming in densely packed suspensions of different particle shapes. Physical Review E 84:1–11. [25] Bitbol, M (1986) Red blood cell orientation in orbit C= 0. Biophys. J. 49:1055–1068. [26] Yao, W et al. (2001) Low viscosity ektacytometry and its validation tested by flow chamber. J. Biomech. 34:1501–1509. 23 [27] Keller, SR, Skalak, R (1982) Motion of a tank-treading ellipsoidal particle in a shear flow. J. Fluid Mech. 120:27–47. [28] Dupire, J, Abkarian, M, Viallat, A (2015) A simple model to understand the effect of membrane shear elasticity and stress-free shape on the motion of red blood cells in shear flow. Soft Matter. [29] Dupont, C, Salsac, AV, Barthès-Biesel, D (2013) Off-plane motion of a prolate capsule in shear flow. J. Fluid Mech. 725:180–198. [30] Cordasco, D, Bagchi, P (2013) Orbital drift of capsules and red blood cells in shear flow. Phys. Fluids 25:091902. [31] Sutera, SP, Mehrjardi, MH (1975) Deformation and fragmentation of human red blood cells in turbulent shear flow. Biophys. J. 15:1–10. [32] Lim, GHW, Wortis, M, R., M (2002) Stomatocyte-discocyte-echinocyte sequence of the human red blood cell: Evidence for the bilayer-couple hypothesis from membrane mechanics. Proc. Natl. Acad. Sci. USA 99:16766–16769. [33] Forsyth, AM, Wan, J, Owrutsky, PD, Abkarian, M, Stone, HA (2011) Multiscale approach to link red blood cell dynamics, shear viscosity, and ATP release. Proc. Natl. Acad. Sci. USA 108:10986–10991. [34] Fåhraeus, R, Lindqvist, T (1931) The viscosity of the blood in narrow capillary tubes. Am. J. Physiol. 96:562–568. [35] Dulinska, I et al. (2006) Stiffness of normal and pathological erythrocytes studied by means of atomic force microscopy. J. Biochem. Biophys. Methods 66:1–11. 24 [36] E. Gibaud. Numerical simulation of red blood cells flowing in a blood analyzer. PhD thesis, Université de Montpellier, 2015. [37] W. R. Dodson III and P. Dimitrakopoulos. Tank-treading of erythrocytes in strong shear flows via a nonstiff cytoskeleton-based continuum computational modeling. Biophys. J., 99:2906–2916, 2010. [38] D. Cordasco, Yazdani, and P. Bagchi. Comparison of erythrocyte dynamics in shear flow under different stress-free configurations. Phys. Fluids, 26(041902), 2014. [39] E. A. Evans and Y. C. Fung. Improved measurements of the erythrocyte geometry. Microvascular Research, 4:335–347, 1972. [40] S. Plimpton. Fast Parallel Algorithms for Short-range Molecular-Dynamics. J. Comp. Phys., 117(1):1–19, 1995. [41] R.D. Groot and P.B. Warren. Dissipative particle dynamics: Bridging the gap be- tween atomistic and mesoscopic simulation. J. Chem. Phys., 107(11):4423–4435, 1997. [42] P. Español and M. Revenga. Smoothed dissipative particle dynamics. Phys. Rev. E, 67:026705, Feb 2003. [43] K. Müller, D. A. Fedosov, and G. Gompper. Smoothed dissipative particle dynamics with angular momentum conservation. J. Comp. Phys., 281:301–315, 2015. [44] D. A. Fedosov, M. Peltomaeki, and G. Gompper. Deformation and dynamics of red blood cells in flow through cylindrical microchannels. Soft Matter, 10(24):4258–4267, 2014. 25 Legends Figure 1: Investigation of RBC shapes at different shear rates in a cone-and-plate rheome- ter. A) Observation of hardened cells by optical (black and white) and confocal (red) microscopy: with increasing γ the formation of highly deformed stomatocytes and then polylobed cells (trilobe and hexalobe) are detected (scale bars are 5 µm). B) Shape dis- tribution of RBCs populations in samples hardened at different shear rates: the three regions in color highlight different regime of decrease of discocytes population. The error bars represent triplicate measurements realized in the two rapidly varying regions. They illustrate the typical variance of the measurements. 26 Figure 2: Microfluidic observations of RBC dynamics in shear flow. A) Time lapse sequence of deformation of RBCs at different γ: Time lapses of 20 ms for the tumbling discocyte, 6 ms for the rolling discocyte, 3.25 ms for the 3 stomatocytes, 0.6 ms for the trilobe in top view, 1.75 ms for the trilobe in side view and 0.6 ms for the hexalobe. Right panel corresponds to analogous time sequences of RBCs obtained with YALES2BIO simulations: Time lapses are given in 1/ γ units: 8 for the tumbling discocyte, 7 for the rolling discocyte, 6 for the stomatocyte and the tumbling deformed stomatocyte, and 7 for the four last cases. B) Stop flow sequences of (left) a trilobed and (right) an hexlobed cell. For the trilobe, the total relaxation time is 1 s and the intermediate images are separated by respectively 0.23 and 0.56 s. For the hexalobe, relaxation occurs in 1.18 s and successive images have a time interval of respectively 0.32 and 0.71 s. (scale bars are 5 µm).. 27 Figure 3: Characterization of RBCs morphology as a function of increasing Ht. A) Shape distributions of HRBCs at 900 s−1 in suspensions with different Ht. For the sake of clarity, the number densities of discocytes and stomatocytes, which are negligible, have been omitted. Inset: black frames on the right, a top view and a cross section of a creased discocyte acquired by confocal microscopy as well as an image of a polylobed irregular shape (multilobe) obtained by optical microscopy (scale bar is 5 µm). Physiological hematocrits in the microcirculation lie in the range 15− 30% indicated by the light yellow region. B) Distributions of RBCs extension for different Ht values measured in SDPD simulations. The extension is defined as the maximum instantaneous cell length in the flow direction, as depicted in one of the insets. The distributions include the data from multiple RBCs in a suspension and are averaged over many time instances. The insets illustrate typical shapes for different values of the cell extension. The two extension regions marked by light yellow and light blue colors schematically depict cell-extension ranges, where polylobed shapes and creased discocytes are observed. 28 Figure 4: Rheology of dense suspensions of RBCs (Ht = 45%). A) Relative viscosity of the suspensions of deformable RBCs in plasma (red) and in PBS/BSA (yellow) as a function of γ in comparison to the suspensions of washed cells hardened at rest (green) and at γ=1500 s−1 (blue). Suspensions with a deformable corpuscular phase show a typical shear-thinning for increasing γ, whereas hardened samples have a nearly Newtonian behavior. SDPD simulation data (black stars) are also shown for deformable cells and agree well with the experimental results for RBCs suspended in plasma or in PBS/BSA. B) Rheology of washed blood in comparison to the suspensions of RBCs in solutions with different dextran concentrations (T=25 ◦C). The effect of viscosity ratio λ is highlighted in the inset by the illustration of single cells flowing in both PBS/BSA and dextran solutions in microfluidics at a comparable shear stress τ. The scale bars are equal to 5 µm. 29
1702.05998
1
1702
2017-02-20T15:09:08
A New Route for the Determination of Protein Structure in Physiological Environment
[ "physics.bio-ph" ]
Revealing the structure of complex biological macromolecules, such as proteins, is an essential step for understanding the chemical mechanisms that determine the diversity of their functions. Synchrotron based x-ray crystallography and cryo-electron microscopy have made major contributions in determining thousands of protein structures even from micro-sized crystals. They suffer from some limitations that have not been overcome, such as radiation damage, the natural inability to crystallize of a number of proteins and experimental conditions for structure determination that are incompatible with the physiological environment. Today the ultrashort and ultra-bright pulses of X-ray free-electron lasers (XFELs) have made attainable the dream to determine protein structure before radiation damage starts to destroy the samples. However, the signal-to-noise ratio remains a great challenge to obtain usable diffraction patterns from a single protein molecule. We describe here a new methodology that should overcome the signal and protein crystallization limits. Using a multidisciplinary approach, we propose to create a two dimensional protein array with defined orientation attached on a self-assembled-monolayer. We develop a literature-based, flexible toolbox capable of assembling different proteins on a functionalized surface while keeping them under physiological conditions during the experiment, using a water-confining graphene cover.
physics.bio-ph
physics
A New Route for the Determination of Protein Structure in Physiological Environment M. Altissimo1*, M. Kiskinova1, R. Mincigrucci1, L. Vaccari1 , C. Guarnaccia2 and C. Masciovecchio1* 1Elettra Sincrotrone Trieste, S. S. 14 km 163 - 34149 Basovizza, Trieste Italy 2International Centre for Genetic Engineering and Biotechnology, Padriciano 99 – 34149 Trieste, Italy Corresponding authors: [email protected], [email protected] Abstract Revealing the structure of complex biological macromolecules, such as proteins, is an essential step for understanding the chemical mechanisms that determine the diversity of their functions. Synchrotron based x-ray crystallography and cryo-electron microscopy have made major contributions in determining thousands of protein structures even from micro-sized crystals. They suffer from some limitations that have not been overcome, such as radiation damage, the natural inability to crystallize of a number of proteins and experimental conditions for structure determination that are incompatible with the physiological environment. Today the ultrashort and ultra-bright pulses of X-ray free-electron lasers (XFELs) have made attainable the dream to determine protein structure before radiation damage starts to destroy the samples. However, the signal-to-noise ratio remains a great challenge to obtain usable diffraction patterns from a single protein molecule. We describe here a new methodology that should overcome the signal and protein crystallization limits. Using a multidisciplinary approach, we propose to create a two dimensional protein array with defined orientation attached on a self- assembled-monolayer. We develop a literature-based, flexible toolbox capable of assembling different proteins on a functionalized surface while keeping them under physiological conditions during the experiment, using a water-confining graphene cover. 1. Introduction The determination of the structure of proteins as well as other macromolecules has historically been prerogative of x-ray crystallography. One of the technique's requirements is the growth of high-quality crystals, which need to be sufficiently large to efficiently diffract x-rays while withstanding radiation damage. This method suffers from two noteworthy constraints, making structure determination extremely difficult or sometimes impossible. The first problem is that many bio-molecules hardly form sufficiently large, high quality crystals or do not crystalize at all. These restrictions are most severe for large protein complexes, such as membrane proteins, which participate in a plethora of fundamental biological processes as receptors, transporters or enzymes, and are therefore responsible of many cellular dysfunctions and potential targets for new therapeutic agents. The second limitation is the unavoidable x-ray radiation damage. Crystal size and radiation damage are inherently linked, since reducing the dose delivered to a single molecule requires large crystals amplifying the signal through Bragg diffraction. Therefore, synchrotron-based experiments are usually performed with cryo-cooled crystals in order to reduce the mass-transport rate due to radiation damage. One of the breakthroughs using the ultrafast and coherent nature of intense X-ray pulses generated by free-electron laser (FEL) is single-shot coherent diffractive imaging performed before the destruction of the object, with diffraction-limited resolution. In life sciences, the ultimate goal for FELs is to solve the structure of large biomolecules without crystallization. This ambitious goal calls for the development of new experimental methodologies1,2. The first attempts to solve the structure of large biomolecules include the development of serial femtosecond crystallography (SFC) at room temperature, as well as the use of jets for continuous delivery of micro/nano crystals into the FEL hit zone. The structure is subsequently determined from the collection of many thousands diffraction patterns3–5. Time-resolved SFC experiments highlighted photo-induced changes in the electronic structure due to charge transfer, encoded in the diffraction pattern of the microcrystals under examination 6,7. Interestingly, serial nano-crystallography has also become feasible with focused beams at synchrotrons8, while microfluidic devices for protein crystallization and on-chip diffraction studies have been developed and demonstrated9. One of the main limitations for a broad applicability of SFC is that thousands of crystals are needed to obtain complete data sets. Considering that only a relatively small fraction of successful hits generates useful diffraction patterns, in order to become a routine analytical tool, SFC still lacks an easy, reliable and inexpensive way of producing a high number of nanocrystals. The limited access to perfect crystals has been an unsolved problem in protein crystallography for decades, further adding to the difficulty in solving the structure of complex bio-molecules. It is interesting to note that some recent diffraction studies performed on XFEL on imperfect crystals have demonstrated that this is not a constraint. The 'neglected' weak continuous scattering arising when the crystals become disordered contains key information to overcome the resolution limits and to solve the tricky phase puzzle 10. As shown in the paper, the continuously modulated background fully encodes the waves diffracted from individual single molecules. Thus, by using coherent diffraction imaging methods, it is possible to achieve real-space images of macromolecules with higher lateral resolution than what is obtainable by ordinary Bragg diffraction. A common feature of all the achievements summarized above is that all the proteins structures solved at synchrotrons or FELs sources are based on the use of 3D crystals. It is interesting to note how, even in the case of 'in-vivo' room temperature SFC experiments, the protein crystal does not represent the protein in its more natural 2D arrangement. Until now 2D-crystallography has almost exclusively been the area of transmission electron microscopy (TEM). Significant progress in protein structure determination and lipid interaction has been made thanks to the use of cryo-TEM and the development of algorithms for recovering amplitude and phase information from recorded TEM images11,12. Given the reduced amount and fixed-target sample delivery in near-native environment, single-shot 2D protein crystallography with FELs was suggested as an attractive alternative13,14 and has been recently demonstrated by proof-of-principle experiments at LCLS. One can also argue that the 2D approach adds the ability to explore protein function and dynamics and it is also an intermediate step towards the extremely ambitious XFELs case - atomic imaging of individual bio-molecules. As in the 3D case, sample preparation is also a crucial step for the 2D case, where fixed-target solutions have to be developed. For 2D cryo-TEM studies, most often the proteins crystals are grown embedded within a lipid bilayer, whereas the first XFEL experiments used dices of silicon nitride windows for harvesting 2D crystals, which are then covered with a thin carbon film. These first experiments indicated that in order to overcome the present resolution limits of 7 Å and truly exploiting the unique XFEL properties, an improvement of both sample preparation and data analysis is crucial. In the present work we put forward a new sample delivery method based on fabricating sample supports for hosting the target protein in a near-native environment. Our approach makes use of patterned silicon nitride membranes, over which a graphene cover ensures the stability of the liquid layer hosting the Protein Of Interest (POI). This is in turn covalently bound onto a chemically modified surface, so that an ordered array of proteins is produced in a hydrated environment and kept as close as possible to physiological conditions during data acquisition. The chemical binding method of the POI to the surface must ensure that the layer is 2D-ordered, and the protein in its biologically active state. Further, we present an experimental approach based on free electron laser diffraction before destroy, the final aim of our work being the determination of protein structure in physiological conditions. 2. Discussion a. Directed protein immobilization The immobilization of a POI on a surface can be accomplished through several methods, both physical and chemical. Although physical methods are in general easier and more straightforward to develop, they tend to yield a disordered array of the POI on the substrate of choice. The added constraint of having and ordered array severely restricts the available choices. The binding process should satisfy the following criteria: 1) It covalently binds two unique and mutually reactive groups, one on the protein and one on the surface; 2) It is bio-orthogonal, i.e. it should not appear in, or cross-react with, any of the groups of endogenous amino acids; 3) It works efficiently under physiological conditions, and without the use of harsh chemicals that could cause the denaturation of the POI; 4) It does not alter substantially the structure of the POI, so as to leave it in its functional state; The added complication is that the structure of each protein varies hugely; this impacts directly on the fact that there is no single method that can be used universally to attach proteins to a surface. Thus, given the POI, the binding methodology should be chosen between a set of several, i.e. a "toolbox". Since the final goal is to determine the structure of a protein, a corollary of points 1) and 4) is that, regardless of the chemistry used to bind the POI to a surface, the reacting moiety should be small in comparison to the POI. Engineering the positioning of proteins on a surface in ordered arrays has been studied extensively by the protein biochip community15,16. The POI will have to be modified with a unique chemical group or sequence at a site-specific location in order to ensure a covalent and oriented binding to the surface, which in turn must also be adapted to "receive" the POI. The adaptation of the POI can be accomplished through its recombinant expression by a suitably modified vector. Several classes of chemical reactions can be employed to engineer proteins onto a surface, and are reviewed in a number of papers15–19. The reactions listed in literature are: a) catalyzed cyclo-addition 20,21, where the ring strain energy of a cyclooctine group catalyzes an azide-alkyne cycloaddition, with the azide group attached to the POI. b) modified Staudinger ligation 22,23 creating a stable amide bond between a POI-side azide group and an ester group attached to the surface. c) Diels-Adler cyclo-addition24,25, where a six-membered unsaturated ring is formed by a pair of dienophile/conjugate diene. d) thiol-ene additions26,27 where a thiol is added to an "-ene" group via a free radical mechanism, usually initiated by UV irradiation; e) oxime ligation28,29, which is the aniline-catalized condensation of an oxyamine or hydrazide with an aldehyde or a ketone, producing a stable oxime linkage. f) Expressed Protein Ligation, such as Split-intein-mediated ligation30,31 or farnesyltransferase- related methods32. While these are not strictly a single chemical reactions, they are used to bind a POI to a surface by means of the naturally-occurring protein trans-splicing process33 (split-intein mediated case), or farnesyltransferase mediated binding19. All these reactions satisfy the 4 requirements listed above, and therefore they constitute an appropriate "toolbox" for anchoring ordered 2-D arrays of proteins on a substrate. Their effectiveness has already been reported34. Among the methods described above, we will concentrate on the split-intein mediated ligation, since it is able to yield a traceless attachment of the POI to a gold surface, with minimal POI modification. The technique has been first described in30. The basic idea described there is to genetically engineer a suitable expression vector, with the most common hosts being E. Coli, P. Pastoris or S. Cerevisiae, to express the POI with an N-intein fragment attached to its C-terminal. On the surface, the complementary C-intein fragment is attached to a modified polyethylene glycol (PEG) linker. When the C-intein and the N-intein fragments interact, they form an active intein domain, binding the POI to the surface. The split intein is naturally spliced into the solution, leaving the POI attached to the surface. Generally speaking, a Self-Assembled Monolayer (SAM) of two differently modified PolyEthylene Glycol (PEG) chains are used, given that they are readily available and chemically well understood35. The shorter of the two is used as a spacer, separating the substrate from the solution, while the second longer one is modified to bind with the POI. b. Proposed methodology The tight integration of sample delivery system and of POI orientation will maximize the chance of a successful protein structure reconstruction. We intend to use the toolbox on different kind of protein in order to demonstrate the potential of our approach. At first, we focus on Human Serum Albumin (HSA), in order to show how the proposed method would allow investigating the structure of a well- characterized system in physiological environment, a condition that is not compatible with standard protein diffraction methods. HSA can been expressed in high yield in the culture medium by the methylotrophic yeast Pichia pastoris both alone or as a fusion protein (see 36–38). The HSA structure has been resolved in 1999 by Sugio and coworkers39; all the relevant protein data can be found at the Protein Data Bank40. HSA has an exposed C-terminal that allows for a N-intein group to be attached through recombinant techniques, see Figure 1. Figure 1: Human Serum Albumine. The backbone of the protein is shown superimposed to a space fill model. The C-terminal of the protein is also indicated. Data is from 39,40. Further, being a protein found in blood, it is safe to assume that the protein will be in its functional conformation in a physiological solution. The protein will be bound on Au, through a tailored SAM. The 10 nm Au layer is going to be deposited on a silicon nitride membrane, and the underlying silicon chip will host a number of spatially separated membranes, so as to line each one up with a single FEL shot and collect the corresponding diffraction pattern. Moving the Si chip to the next window, and repeating the process on another membrane will allow to collect the diffraction pattern in a step-and- repeat fashion. A graphene cover will prevent the evaporation of the physiological solution in the vacuum chamber41. This would be ideal, due to the transparency of the graphene layer to FEL photons. As an alternative, a second set of nitride membranes could be used for sealing the wet chamber. Standard photolithography techniques will be used to open silicon nitride membranes and define 10 nm thick, 10 µm in side Au square patches, using 2 nm Ti as an adhesion layer. The periodicity of the FEL- radiation transparent areas on the substrate is determined by the thickness of the Si substrate, due to the angle between the Si (111) and (100) planes. In order to provide the graphene cover with a set of supports, an optical resist will be patterned via photolithography to "corral" each of the membranes. The photoresist props also serve to keep the POI environment wet and to create separated micro- environments, in order to minimize the possibility of water leakage after each FEL shot. Figure 2 schematizes the end result, excluding the graphene cover. ! ! ! !! ! ! ! ! Figure 2: schematics of a single FEL-transparent area. The silicon nitride membrane (in light blue) is used as a carrier for the 10 micron wide, 10 nm thick Au square (in gold). Outside of the membrane perimeter, the area is corralled by patterned photoresist, used as a prop for the graphene cover (not shown in the figure) and to keep each sample area wet, and separated from the others. To define the SAM, we will proceed as described in 19 with a slight modification. A dithiocarbamate (DTC)-modified PEG will serve as a linker, with the loose end (i.e. the one not attached to the underlying gold layer) modified by a C-intein domain as described in 16,30. A shorter, unmodified DTC- PEG layer will be used as a spacer. The density of the anchoring points can be controlled by adjusting the relative concentration of the two DTC's in solution. The POI (specifically HSA) will be recombinantly over-expressed by genetically engineered Pichia Pastoris42 or by other suitable eukaryotic vector. One of the key points of the binding process proposed here is that the POI will bind to the surface directly from the cell culture medium thus avoiding lengthy purification steps. The POI will be covalently linked to the Au surface via the intein-mediated trans-splicing process as described C ! ! N"intein C"intein ! ! !!!!!!!!!!! ! ! ! ! !!!!!!!!!! functional intein ! Self"spliced Split&intein ! !!!!!!!!!!! !! above. Figure 3: schematics of HSA covalent binding process to the surface. All the other cell debris will be washed away with an appropriate rinsing step. Figure 3 shows schematically the binding process. By rough calculation, according to the data published on the Protein Data Bank, we estimate that there will be approximately 2.5 million HSA proteins per 10 µm square Au patch. Once the proteins have been attached to the surface, a graphene layer will be added on top to seal the stack. What we believe is particularly interesting about our methodology is the fact that any protein of known sequence can potentially be recombinantly modified to incorporate into either the N- or the C-terminal a suitable moiety that will bind to a chemically complementary substrate, regardless of the ability of the POI to crystallize. As stated above, the density of the surface anchoring points will be determined by the relative concentration of the spacer/linker PEG chains. Coupled to this, the protein-protein interactions due to charge distribution and steric hindrance will order of the POI onto the chemically-engineered surface. This has been demonstrated in literature by, for example19,43. c. Experimental Needs In this section we will discuss the experimental requirements needed to reconstruct the protein structure. To avoid sample degradation due to radiation damage, FELs pulses represent a valuable choice44. Accordingly to45, diffraction images can be easily acquired to better than 8.5 Å resolution, from two-dimensional protein crystal samples less than 10 nm thick and at room temperature. While this achievement represents a big step in the direction of determining protein structure in physiological condition, the limits of the approach proposed in 45 were basically related to the fact that the protein had random orientation on the substrate. As the authors pointed out: "The extent to which multiple lattice or powder diffraction XFEL data can be reliably used for structure determination from 2-D crystals is currently unclear". Using our approach in the case of HSA we can align up to about 2.5 · 108 proteins on a silicon nitride substrate as large as 100 mm2. Figure 4 – Proposed set-up for the single beam diffraction (top left) and for double beam diffraction (top right). The angle of incidence of the beam with respect to the surface and/or the inter-beam angle can be varied to obtain complementary diffraction patterns. This will strongly increase the intensity from Bragg diffraction and, moreover, the use of a large focal spot will prevent electron stripping from the atoms upon pulse arrival, thus decreasing the scattering from unbound electrons and further increasing the signal-to-noise ratio. Finally, the use of a graphene cover allows keeping the protein in physiological condition. Focal spot sizes can be easily varied from 10 µm to 1 mm using standard optics set-ups, thus defining the maximum POI island size that can be illuminated by a single shot. In case of ablation, samples can be raster scanned. The diffraction pattern of each POI patch may be acquired with a suitable in-vacuum area detector. When single shot techniques have been applied to nano-crystals, the 3D structure of the protein has been recovered by sorting and merging diffraction patterns obtained by randomly oriented crystals in liquid jets3. With our approach, one can rotate the sample under the beam so as to precisely define the angle under which the POI is observed. Alternatively, a set-up consisting of two crossed FEL beams46 (Figure 4) and two detectors may be employed to simultaneously record the diffraction pattern of the same area from two different directions. Figure 5- Estimated sample transmission for 50 nm of Silicon Nitride and 0.5 mm of water (blue curve) and for 100 nm of Silicon Nitride and 2 mm of water (orange curve). Figure 5 shows the calculated theoretical transmission of the sample stack. The blue and the orange curve refer to 50 nm nitride/0.5 µm water, and 100 nm nitride/1 µm water respectively. For both of the curves, a 2 nm Ti and 10 nm Au bilayer has been considered. As expected, a monolayer of proteins does not contribute substantially to the total absorption of the stack. The proposed structure has a static transmission value around 40 % in the least favorable case for photon energies above 1 keV. In order to calculate the signal improvement caused by the orientation of the POI, it is interesting to estimate the increase in the Bragg signal due to the number of proteins aligned on the surface and under the beam. As in any interference problem, the strength of the main peak is proportional to the number of diffracting objects squared47, while the ratio between different order/directions is governed by more complex relations48. For photosystem I as studied by3 with single FEL shot methods, an intensity in the Bragg spots ranging from 5 x 103 to 5 photons was recorded as function of angle. The lowest reported crystal size (P63 crystal structure with a diameter of 160 nm and a diameter to length ratio of 1:2) corresponds to a number of proteins ranging from 270 to 540. With our sample delivery method, used for the same photosystem I, we expect to obtain a Bragg peak intensity which is at least a factor 6 x 106 higher, since a POI SAM of 100 x 100 mm should contain ~ 1.3 x 107 proteins. Similar ratios are thus expected for others proteins. The recent development of FEL Facilities has opened up different schemes for the production of photon pulses49. One recently highlighted possibility is to allow only a single superradiant spike of the electron bunch self-emission to be amplified in an exotic configuration. This mode may produce pulses as short as 500 attosecond, with the drawback of a limited number of photons per pulse. Using the parameters of the European XFEL under commissioning in Hamburg50, the expected flux would be about 1011 photon per pulse. It has been demonstrated that FEL single molecule imaging will strongly benefit from using sub-femtosecond pulses, due to the significant radiation damage and the formation of preferred multisoliton clusters, which reshape the overall electronic density of the molecular system when using longer pulses51. While the expected number of photons per pulse in the superradiance scheme would be too low to get the diffraction pattern from a single molecule, this is not the case for the 2D array sample delivery method we introduce here, thanks to the several orders of magnitude enhancement of the scattering power due to the large number of molecules involved in the process. Conclusions 3. In this article we propose a new method capable of delivering two-dimensional, aligned arrays of bio- molecules. The alignment should increase the signal-to-noise ratio by several orders of magnitude on FEL-based diffraction experiments, as compared with other reported studies. The sample will be a two- dimensional crystal containing up to billions of ordered macromolecules that will be kept in a near- native environment. The route presented above follows a multidisciplinary approach based on existing literature. We discussed in detail the advantages of this methodology that we highlight in the following: 1) It will permit to reveal the structure of macromolecules in general, and proteins more in particular, that cannot be crystallized and, therefore, cannot be studied by classical crystallography; 2) The 2D protein array would allow to use FEL pulses produced in exotic configurations with shorter pulse duration (namely few attosecond). This would definitively avoid sample radiation damage during the diffraction process. 3) The measurements can be carried out under physiological conditions; 4) Only a moderate sample quantity will be necessary. This is extremely important when samples are difficult or very expensive to produce. Moreover, our approach avoids lengthy purification steps of the POI, due to the bio-orthogonal chemical reaction binding it to the engineered surface. It is worthwhile noting that a different ordering strategy is required in order to solve the structure of membrane proteins in their native environment. These constitute 20-30% of the total proteome, but represent about 1% of determined protein structures, due to difficulties in their crystallization. Cryo- TEM has been one of the main tools of choice for integral membrane proteins (IMP) structure reconstruction. The base of the 2D crystallization behind cryo-TEM IMP structure reconstruction is a work recognizing as early as 199252 that IMP can form 2D crystals when inserted into an artificial phospholipid bilayer 53,54. Due to the nature of IMPs, ad-hoc crystallization conditions have to be determined for each protein55–57. A toolbox for the 2D crystallization of IMPs is therefore already well established, and we suggest that FEL-based in particular, and more in general x-ray, IMP structure reconstruction should follow the methodologies found in literature. An engineered sample delivery method can be used to exploit the two main characteristics of FEL pulses, i.e. tunability and coherence. The former can be used to perform anomalous scattering experiments across the edge of biologically abundant elements, such as S and P. The coherence of the beam, coupled with our sample holder, will allow for precisely reconstructing the phase of the scattered photons. We note finally that the sample delivery method we propose here allows for pump-and-probe experiments, whereby a visible laser excites the POI, and the FEL beam probes its structure post- excitation. This could, for example, enable to determine the dynamic change in the structure of a photosynthetic centre, before, during and after its excitation by visible-light. 4. Acknowledgments The authors acknowledge G. Kourousias for useful discussions on phase retrieval methods. 5. References 1. 2. 3. 4. 5. 6. 7. 8. 9. Neutze, R., Brändén, G. & Schertler, G. F. X. Membrane protein structural biology using X-ray free electron lasers. Curr. Opin. Struct. Biol. 33, 115–125 (2015). Kern, J., Yachandra, V. K. & Yano, J. Metalloprotein structures at ambient conditions and in real-time: Biological crystallography and spectroscopy using X-ray free electron lasers. Curr. Opin. Struct. Biol. 34, 87–98 (2015). Chapman, H. N. et al. Femtosecond X-ray protein nanocrystallography. Nature 470, 73–77 (2011). Gallat, F.-X. et al. In vivo crystallography at X-ray free-electron lasers: the next generation of structural biology? Philos. Trans. R. Soc. B Biol. Sci. 369, 20130497–20130497 (2014). Schlichting, I. Serial femtosecond crystallography: The first five years. IUCrJ 2, 246–255 (2015). Kern, J. et al. NIH Public Access. 340, 491–495 (2013). Kupitz, C. et al. Serial time-resolved crystallography of photosystem II using a femtosecond X- ray laser. Nature 513, 261–265 (2014). Gati, C. et al. Serial crystallography on in vivo grown microcrystals using synchrotron radiation. IUCrJ 1, 87–94 (2014). Heymann, M. et al. Room-temperature serial crystallography using a kinetically optimized microfluidic device for protein crystallization and on-chip X-ray diffraction. IUCrJ 1, 349–360 (2014). 10. Ayyer, K. et al. Macromolecular diffractive imaging using imperfect crystals. Nature 530, 202– 206 (2016). 11. Hite, R. K., Raunser, S. & Walz, T. Revival of electron crystallography. Curr. Opin. Struct. Biol. 17, 389–395 (2007). 12. Gipson, B. R. et al. Automatic recovery of missing amplitudes and phases in tilt-limited electron crystallography of two-dimensional crystals. Phys. Rev. E - Stat. Nonlinear, Soft Matter Phys. 84, 1–9 (2011). 13. Kewish, C. M., Thibault, P., Bunk, O. & Pfeiffer, F. The potential for two-dimensional crystallography of membrane proteins at future x-ray free-electron laser sources. New J. Phys. 12, (2010). 14. 15. Proteomics, C. Architecture. Steen Redeker, E. et al. Protein engineering for directed immobilization. Bioconjug. Chem. 24, 1761–1777 (2013). 16. Camarero, J. A. Recent developments in the site-specific immobilization of proteins onto solid supports. Biopolym. - Pept. Sci. Sect. 90, 450–458 (2008). 17. Rusmini, F., Zhong, Z. & Feijen, J. Protein immobilization strategies for protein biochips. Biomacromolecules 8, 1775–1789 (2007). 18. Chen, Y. X., Triola, G. & Waldmann, H. Bioorthogonal chemistry for site-specific labeling and surface immobilization of proteins. Acc. Chem. Res. 44, 762–773 (2011). 19. Choi, S. R., Seo, J. S., Bohaty, R. F. H. & Poulter, C. D. Regio- and chemoselective immobilization of proteins on gold surfaces. Bioconjug. Chem. 25, 269–275 (2014). 20. Baskin, J. M. et al. Copper-free click chemistry for dynamic in vivo imaging. Proc. Natl. Acad. Sci. U. S. A. 104, 16793–7 (2007). 21. Agard, N. J., Prescher, J. A. & Bertozzi, C. R. A strain-promoted [3 + 2] azide-alkyne cycloaddition for covalent modification of biomolecules in living systems. J. Am. Chem. Soc. 126, 15046–15047 (2004). 22. 23. Staudinger, H. & Hauser, E. ??ber neue organische Phosphorverbindungen IV Phosphinimine. Helv. Chim. Acta 4, 861–886 (1921). Saxon, E. & Bertozzi, C. R. Cell Surface Engineering by a Modified Staudinger Reaction. Science (80-. ). 287, 2007–2010 (2000). 24. Houseman, B. T., Huh, J. H., Kron, S. J. & Mrksich, M. Peptide chips for the quantitative evaluation of protein kinase activity. Nat. Biotechnol. 20, 270–274 (2002). 25. De Araújo, A. D. et al. Diels-Alder ligation and surface immobilization of proteins. Angew. Chemie - Int. Ed. 45, 296–301 (2005). 26. Poster, T. No Title. Ber. Deutsch. Chem. Ges. 38, 646−657 (1905). 27. Hoyle, C. E. & Bowman, C. N. Thiol-ene click chemistry. Angew. Chemie - Int. Ed. 49, 1540– 1573 (2010). 28. Kaumaya, P. Protein engineering. (InTech, 2012). 29. Dirksen, A. & Dawson, P. E. Rapid oxime and hydrazone ligations with aromatic aldehydes for biomolecular labeling. Bioconjug. Chem. 19, 2543–2548 (2008). 30. Kwon, Y., Coleman, M. A. & Camarero, J. A. Selective immobilization of proteins onto solid supports through split-intein-mediated protein trans-splicing. Angew. Chemie - Int. Ed. 45, 1726–1729 (2006). 31. Braner, M., Kollmannsperger, A., Wieneke, R. & Tampé, R. 'Traceless' tracing of proteins – high-affinity trans-splicing directed by a minimal interaction pair. Chem. Sci. 2646–2652 (2016). doi:10.1039/C5SC02936H 32. Gauchet, C., Labadie, G. R. & Poulter, C. D. Regio- and chemoselective covalent immobilization of proteins through unnatural amino acids. J. Am. Chem. Soc. 128, 9274–9275 (2006). 33. Wu, H., Hu, Z. & Liu, X. Q. Protein trans-splicing by a split intein encoded in a split DnaE gene of Synechocystis sp. PCC6803. Proc Natl Acad Sci U S A 95, 9226–9231 (1998). 34. Lin, P. C., Weinrich, D. & Waldmann, H. Protein biochips: Oriented surface immobilization of proteins. Macromol. Chem. Phys. 211, 136–144 (2010). 35. Samanta, D. & Sarkar, A. Immobilization of bio-macromolecules on self-assembled monolayers: methods and sensor applications. Chem. Soc. Rev. 40, 2567–2592 (2011). 36. Zhu, R.-Y. et al. Expression and purification of recombinant human serum albumin fusion protein with VEGF165b in Pichia pastoris. Protein Expr. Purif. 85, 32–7 (2012). 37. Kobayashi, K. et al. High-level expression of recombinant human serum albumin from the methylotrophic yeast Pichia pastoris with minimal protease production and activation. J. Biosci. Bioeng. 89, 55–61 (2000). 38. Belew, M., Yan, M. & Caldwell, K. Purification of Recombinant Human Serum Albumin (rHSA) Produced by Genetically Modified Pichia Pastoris. Sep. Sci. Technol. 43, 3134–3153 (2008). 39. Sugio, S., Kashima, a, Mochizuki, S., Noda, M. & Kobayashi, K. Crystal structure of human serum albumin at 2.5 A resolution. Protein Eng. 12, 439–446 (1999). 40. Protein Data Bank. Available at: http://www.rcsb.org/pdb/explore/explore.do?structureId=1AO6. 41. Berry, V. Impermeability of graphene and its applications. Carbon N. Y. 62, 1–10 (2013). 42. Chuang, V. T. G. and Otagiri, M. Recombinant human serum albumin. Drugs of Today 8, 547 (2007). 43. Weinrich, D. et al. Oriented immobilization of farnesylated proteins by the thiol-ene reaction. Angew. Chemie - Int. Ed. 49, 1252–1257 (2010). 44. Neutze, R., Wouts, R., van der Spoel, D., Weckert, E. & Hajdu, J. Potential for biomolecular imaging with femtosecond X-ray pulses. Nature 406, 752–757 (2000). 45. Frank, M. et al. Femtosecond X-ray diffraction from two-dimensional protein crystals. IUCrJ 1, 95–100 (2014). 46. Bencivenga, F. et al. Four-wave mixing experiments with extreme ultraviolet transient gratings. Nature 520, 205–8 (2015). 47. Jackson, J. D. Classical electrodynamics. (1962). 48. Ashcroft, N. W. & Mermin, David, N. Solid State Physics. (1976). 49. Bonifacio, R., Fares, H., Ferrario, M., W. J. McNeil, B. & R. M. Robb, G. Design of sub- Angstrom compact free-electron laser source. Opt. Commun. 382, 58–63 (2017). 50. The Technical Design Report (TDR) of the European XFEL. Available at: https://xfel.desy.de/technical_information/tdr/tdr/. 51. Fratalocchi, A. & Ruocco, G. Single-molecule imaging with X-ray free-Electron Lasers: Dream or Reality? Phys. Rev. Lett. 106, 1–4 (2011). 52. Kühlbrandt, W. Two-dimensional crystallization of membrane proteins. Q. Rev. Biophys. 25, 1– 49 (1992). 53. Coutable, A. et al. Preparation of tethered-lipid bilayers on gold surfaces for the incorporation of integral membrane proteins synthesized by cell-free expression. Langmuir 30, 3132–3141 (2014). 54. Giess, F., Friedrich, M. G., Heberle, J., Naumann, R. L. & Knoll, W. The Protein-Tethered Lipid Bilayer: A Novel Mimic of the Biological Membrane. Biophys. J. 87, 3213–20 (2004). 55. De Zorzi, R. et al. Growth of large and highly ordered 2D crystals of a K+ channel, structural role of lipidic environment. Biophys. J. 105, 398–408 (2013). 56. Signorell, G. A., Kaufmann, T. C., Kukulski, W., Engel, A. & Rémigy, H. W. Controlled 2D crystallization of membrane proteins using methyl-??-cyclodextrin. J. Struct. Biol. 157, 321–328 (2007). 57. Ubarretxena-Belandia, Iban, and Stokes, D. L. Recent Advances in Electron Cryomicroscopy. Recent Advances in Electron Cryomicroscopy, Part A 81, (2010).
1509.06450
2
1509
2015-09-23T03:50:35
Flexibility of nucleic acids: from DNA to RNA
[ "physics.bio-ph", "cond-mat.soft", "cond-mat.stat-mech", "physics.chem-ph" ]
The structural flexibility of nucleic acids plays a key role in many fundamental life processes, such as gene replication and expression, DNA-protein recognition, and gene regulation. To obtain a thorough understanding of nucleic acid flexibility, extensive studies have been performed using various experimental methods and theoretical models. In this review, we will introduce the progress that has been made in understanding the flexibility of nucleic acids including DNAs and RNAs, and will emphasize the experimental findings and the effects of salt, temperature, and sequence. Finally, we will discuss the major unanswered questions in understanding the flexibility of nucleic acids.
physics.bio-ph
physics
Flexibility of nucleic acids: from DNA to RNA Bao Lei (鲍磊), Zhang Xi (张曦), Jin Lei (金雷), and Tan Zhi-Jie (谭志杰)* Department of Physics and Key Laboratory of Artificial Micro- and Nano-structures of Ministry of Education, School of Physics and Technology, Wuhan University, Wuhan 430072, China Abstract The structural flexibility of nucleic acids plays a key role in many fundamental life processes, such as gene replication and expression, DNA-protein recognition, and gene regulation. To obtain a thorough understanding of nucleic acid flexibility, extensive studies have been performed using various experimental methods and theoretical models. In this review, we will introduce the progress that has been made in understanding the flexibility of nucleic acids including DNAs and RNAs, and will emphasize the experimental findings and the effects of salt, temperature, and sequence. Finally, we will discuss the major unanswered questions in understanding the flexibility of nucleic acids. PACS: 87.10.Pq, 87.14.G-, 87.15.ap, 87.15.H-, 87.15.La Key words: DNA, RNA, structure, flexibility, salt, temperature, sequence, computer simulation _____________________________________________________________________ *Corresponding author: [email protected] 1 1. Introduction Nucleic acids are negatively charged biopolymers and their structures are generally stabilized by base pairing/stacking interactions and metal ion-binding.[1−4] Due to the polymeric nature of nucleic acids and the stabilizing energy on the order of ~kBT (thermal energy), nucleic acids generally show strong conformational fluctuations and are rather flexible. The flexibility of nucleic acids is extremely important for their biological functions such as gene replication and expression, protein recognition, and gene regulation.[1−5] To evaluate the flexibility of nucleic acids, various experimental methods have been developed, such as atomic force microscopy (AFM),[6,7] fluorescence resonance energy transfer (FRET),[8,9] small-angle x-ray scattering (SAXS),[9−11] magnetic tweezers (MT),[12−14] and optical tweezers (OT),[14−16] among others. Some theoretical models have also been developed and combined with experimental approaches to quantify the flexibility of nucleic acids, such as the worm-like chain (WLC) model and free-joint chain model.[17,18] Recently, with the development of computational methods, molecular dynamics (MD) simulations[19−24] have been widely used to examine the flexibility of nucleic acids at the atomic level. Beyond the experiments that can only evaluate the macroscopic properties of flexibility, MD simulations and theoretical modeling at the atomic level can reveal detailed microscopic information, such as conformational changes of nucleic acids and ion binding patterns, as well as related microscopic mechanisms.[21,24] The use of various advanced experimental techniques and atomistic MD/modeling has greatly enhanced the understanding of nucleic acid flexibility.[6−22] Due to the order of ~kBT for base pairing/stacking energy and the polyanionic nature of nucleic acids,[1−5,25] their flexibilities strongly depend on their sequences, salt ions in solution, and temperature, which would affect the strength of base pairing/stacking, ion binding, and chain conformational entropy, respectively. Drugs or proteins can also interact with nucleic acids and dramatically affect their structures and flexibilities.[2,3,5,26] In addition, nucleic acids can exhibit distinctively different flexibilities for different structural states depending on the temperature and ionic conditions.[27−31] Therefore, nucleic acid flexibility is influenced by several critical factors, such as sequence, salt, and temperature. In this review, we will focus on recent progress made in understanding the flexibility of nucleic acids. Since the family of nucleic acids includes single-stranded (ss) DNA/RNA, double-stranded (ds) DNA/RNA, and a large number of RNA tertiary folds, the main text is organized as follows. First, we will provide a brief overview of the flexibility of ssDNA/RNA. Second, we will focus on the flexibility of dsDNA, which has already attracted much attention 2 for many years. Third, we will describe recent progress in understanding the flexibility of RNAs. Finally, we will discuss the major unanswered questions in understanding the flexibility of nucleic acids. 2. Flexibility of ssDNA and ssRNA The ss chain is an elementary structural and functional segment of nucleic acids. For example, RNA structures generally consist of different types of loops, and the ss chain also represents the denatured state of nucleic acids.[1,32,33] Furthermore, the ss chain is an important intermediate in many key biochemical processes, such as replication, recombination repair, and transcription, and is specifically recognized by many proteins.[32,34] The flexibility of the ss chain plays a significant role in its interactions with other macromolecules, such as proteins.[35] Generally, under the physiological conditions, ss nucleic acid chains composed of generic sequences are rather flexible, and can be approximately described using the free-joint chain model, while ss nucleic acid flexibility may be sensitive to the sequence and ionic environment.[36−52] The flexibilities of ss nucleic acids have been quantified using various experimental approaches, such as force-extension curves,[36−40] FRET,[41,42] SAXS,[43] AFM,[44] and fluorescence recovery after photobleaching.[52] These experimental measurements are summarized in Table 1. In addition, computer simulations, such as atomistic molecular and Monte Carlo simulations, have been employed to evaluate the flexibility of ss nucleic acids, including the effects of sequence and salt.[36−52] 2.1. Sequence effect Various experiments have suggested that the structure and flexibility of an ss DNA/RNA chain strongly depends on the intra-chain interactions, such as base-pairing and base stacking, which are highly correlated with the nucleic acid sequence. An ssDNA chain composed of a generic sequence can fold into a secondary structure, such as a hairpin or a helix through base-pairing/stacking and different sequence compositions give generic ss chains different properties, including flexibility.[36,40,45] Single-molecule experiments have suggested that an ssDNA can be modeled as the free-joint-chain model under physiological ionic conditions with a small persistence length of ~7.5 Å,[49] since strong intra-chain base stacking is generally not observed in these molecules. A homo-polynucleotide ss DNA may exhibit self-stacking interactions between nearby nucleotide bases, which can be sufficiently strong to maintain the rigidity of the ss chain.[44] The 3 varying strengths of intra-chain self-stacking interactions results in different flexibilities of the four types of homo-polynucleotide ss DNA molecules, including poly(dA), poly(dT), poly(dC), and poly(dG). Previous experiments have shown that poly(dT) is much more flexible than poly(dC), poly(dG), and poly(dA), since the self-stacking interactions in these molecules were much stronger than that in poly(dT). Actually, poly(dT) behaves as a free polyelectrolyte chain,[39,44,50] while poly(A), poly(G), and poly(C) would form ordered ss helices because of strong intra-chain self-stacking.[33,39] Since thymine (T) and uridine (U) show similar strengths in base-stacking interactions, the effect of sequence on the flexibility of ssRNA is very similar to that of DNA.[39,51] A previous study showed that the intra-chain self-stacking of ssRNA is slightly stronger than that of ssDNA because of the difference in the sugar ring. Slightly lower flexibility for poly(rU) than for poly(dT) has been observed under the same ionic conditions.[47] However, the details of self-stacking interactions in ssDNA/RNA and the mechanism of the flexibility of ssDNA/RNA remain unclear. Advanced experimental techniques and improved force fields for atomistic simulations can be used to examine these issues. 2.2. Effects of salt and chain length Due to the similar charge density on ssDNA and ssRNA, the ion dependence of ss RNA flexibility is very similar to that of ssDNA.[41,47,53] Ions in solution can bind to ssDNA/RNA, which would increase nucleic acid flexibility by neutralizing the negative charges on the phosphates. Numerous experiments have shown that ssDNA/RNA becomes more flexible with increasing concentrations of ions, including Na+ and Mg2+, and this ion-dependence of flexibility is stronger in longer sequences.[36,38,41−43,47,48] Mg2+ has a higher ionic charge than Na+, and experiments and simulations have shown that Mg2+ is approximately 60−120-fold more efficient than Na+ in ionic neutralization.[41,47] Additionally, the ion concentration dependence of ss DNA/RNA flexibility is stronger for Na+ than for Mg2+.[36,38,41,47] At high salt concentration, intra-sequence self-stacking interactions would dominate the global flexibility.[42,47] For ss generic sequences, empirical formulas have been derived for ion-dependent persistence length, which are practically useful.[43,47,48] For an ssDNA under force, a previously derived formula can describe the force-dependent ssDNA force-extension curves at various concentrations of NaCl.[54] Previous studies have also shown that the ion-dependence of ssDNA/RNA flexibility depends on sequence length, i.e., the persistence length P of longer ssDNA/RNAs increases more rapidly than that of short sequences when the ion concentration is decreased.[36,38,41,47,53] This 4 occurs because ion binding to ssDNA/RNAs strongly depends on sequence length.[36,38,41−43,47,52] Experiments have also shown that there is scaling law between the size of ssDNA and its length, and the scaling exponent decreases with increasing monovalent salt concentration.[55] However, the effect of multivalent ions on ssDNA/RNA flexibility remains unclear since the strong ion-ion correlation and possible intra-chain base-pairing/stacking can become tightly coupled. 3. Flexibility of dsDNA 3.1. General features in flexibility of dsDNA Since the discovery of the dsDNA helix, numerous studies have demonstrated that dsDNAs are extensively involved in various life processes.[1,2,56,57] For example, dsDNA can fold into compact structures to enter into bacteriophage heads[56] or form chromatin,[57] and dsDNA can also sharply bend on a local scale to execute its biological functions, such as replication, DNA repair, and transcription, among other functions.[1,2] Therefore, understanding the biological processes related to dsDNAs requires the comprehensive understanding of dsDNA flexibility. DsDNA, as a highly dynamic structure, is stretchable, bendable, and twistable in vivo and in vitro, and its flexibility can be characterized by the three important elastic parameters: stretching modulus S, bending persistence length P, and torsional persistence length C. S, P, and C describe the stretching, bending, and twisting flexibilities, respectively. In addition, contour length (L), end-to-end distance (Ree), and radius of gyration (Rg) have also been used to describe the global structural flexibility of dsDNAs. Extensive experiments have been conducted to quantify the flexibility of dsDNA, and the experimental measurements are summarized in Table 2. Previous reviews[1,2] have described the bending persistence length P of long dsDNA in buffers of moderate salt concentration, which was 45−50 nm based on early experiments. Recently, advanced single-molecule techniques have enabled accurate measurements of dsDNA flexibility. Herrero-Galan et al. manipulated long dsDNAs using magnetic tweezers and optical tweezers, and they observed a bending persistence length P of ~49 ± 2 nm and a stretch modulus S of ~935 ± 121 pN at 150 mM NaCl.[15] Other recent force-extension experiments for dsDNA, indicated that torsional persistence length C was 100 ± 7 nm,[58] a slightly higher value than is generally accepted.[59−62] Very recently, Dekker et al. explored the elastic properties of dsDNA and derived all four elastic constants for dsDNA at 100 mM monovalent salt: 45 ± 2 nm for P, 1000 ± 200 pN for S, 109 ± 4 nm for C, and a negative twist-stretch coupling parameter of ~17 ± 5.[14] Since dsDNA is generally stabilized by specific base-pairing/stacking interactions and the 5 binding of metal ions, the flexibility of dsDNA is strongly dependent on salt, sequence, and temperature. Next, we will comprehensively summarize the flexibility of dsDNAs in the three aspects. Finally, we will introduce recent findings for short dsDNA and dsDNA under high force. 3.2. Salt effect Due to the highly negative charges on dsDNA, the flexibility and stability of dsDNA are tightly coupled to the metal ions present in solution.[1,28,63,64] Numerous experimental and theoretical studies have focused on the role of salt in dsDNA flexibility and have revealed the following major features:[15,27,28,65−74] 1. The increase of monovalent salt concentration enhances the flexibility of dsDNA, which is reflected by the decrease in bending persistence length P. P can decrease to ~45−50 nm at high (~1 M) salt concentration.[15,27,28] 2. The addition of multivalent (≥ 2+) salt clearly enhances the flexibility of dsDNA[63] and causes the collapse of long dsDNAs into compact condensates, which is reflected by a persistence length (~20−40 nm) much lower than ~50 nm.[27,28,74] 3. The stretching modulus S is also strongly dependent on salt. Increasing salt concentration increases S, and S is larger for higher valent salts.[15,28] 4. With increasing monovalent salt concentration, the contour length and twist angle both decrease linearly as a logarithmic function of salt concentration.[75,76] However, on the salt effect on dsDNA flexibility, there are still important questions remained. First, the strength of the salt dependence of P of dsDNA, particularly above ~0.1 M monovalent salt, is unclear. Odijk, Skolnick and Fixman have previously proposed the OSF theory, which divides bending persistence length into intrinsic and electrostatic contributions.[77,78] OSF theory predicts that the electrostatic part only contributes less than ~10% to the total persistence length of dsDNA under physiological ionic conditions. Nevertheless, Manning developed a model based on his counterion condensation theory by defining a hypothetical structure of dsDNA in the absence of DNA residual charges as a “null isomer of DNA.”[79] Manning’s model predicted that the dependence of P on salt concentration is nearly log-[salt] over the entire concentration range and the electrostatic contribution to total bending persistence length can reach ~86%, which has been supported by several experiments.[67−69] Savelyev et al. developed a two-bead coarse-grained structural model for dsDNA and conducted a reexamination by MD simulations. They found that both electrostatic and nonelectrostatic interactions play comparable roles in dsDNA flexibility.[80] They also found that dsDNA bending persistence length decreases by ~25% when monovalent salt is decreased from 0.1 M to 1 M.[81] 6 Therefore, there is still no consensus regarding how DNA flexibility depends on salt concentration, particularly at high salt concentrations of > 1 M. Second, how multivalent ions influence the flexibility of dsDNA is still unclear since multivalent ions interact with DNA much more strongly than monovalent ions.[70,71] As indicated in experiments, multivalent ions such as Mg2+ and Co(NH3)6 3+, have very different effects on DNA flexibility at the same ionic strength compared with monovalent ions. For example, 3+ can lead to a value of P for dsDNA as low as 25−30 nm.[28] Is the effect of Co(NH3)6 multivalent ions coupled to ion-mediated effective interaction between segments in dsDNA?[82,83] Furthermore, what are the roles of divalent ions in the flexibility of dsDNA since some divalent ions can induce effective attractions between dsDNA helices, while other divalent ions cannot?[84] In addition to the unsolved issues described above, additional questions remain at the microscopic level: (1) Why does the stretching modulus increase with increasing ion concentration and ion valence? (2) Why are stretching and twisting negatively coupled? 3.3. Temperature effect For a polymer, the WLC model describes the relationship between bending rigidity g and bending persistence length P as:[85] , (1) where kB is the Boltzmann constant and T is absolute temperature. If g does not change with temperature, the temperature-dependence of P can be described by Eq. (1), and P decreases linearly with increasing temperature. However, dsDNA is composed of sequential base pairs unlike an ideal polymer. Since base pairing/stacking strength is on the order of ~kBT, temperature plays an important role in dsDNA flexibility. With increasing temperature, bases and backbone of dsDNA fluctuate more strongly, which may cause local “melted” bubbles. If temperature becomes sufficiently high, dsDNA strands can become completely separated and exhibit a “melted” state of ss chain. Therefore, the actual temperature dependence of P should be much stronger than that predicted from Eq. (1). Approximately 40 years ago, Gray and Hearst measured the sedimentation coefficient of DNA at infinite dilution and obtained the temperature dependence of P.[29] However, their data showed that the temperature-dependence of P is even weaker than the value predicted from Eq. (1), which assumes the temperature-independent bending rigidity g. Recently, dsDNA cyclization experiments have indicated that local melting in dsDNA can enhance the flexibility of dsDNA 7 BgPkT more significantly than that predicted from the WLC model in the temperature range of 23−42C.[30] A significant decrease in apparent bending persistence length may result from potential excitations of flexible defects,[86,87] which can be excited by high temperature.[88] In order to more accurately measure the temperature dependence of persistence length, Geggier and Vologodskii employed two different approaches to determine the persistence lengths of dsDNA at different temperatures, and the data from the two independent approaches were highly consistent. The experiments showed that the bending persistence length decreased nearly linearly from 53 nm to 44 nm as temperature decreased from 5C to 42C, but decreased very sharply at higher temperatures, reaching ~36 nm at 60C.[31] Corresponding theoretical modeling based on the Peyrard-Bishop-Dauxois model has also shown good agreement with the experimental data. Therefore, both of the experiments by Vologodskii et al. and the Peyrard-Bishop-Dauxois-based model showed the discrepancies with the WLC model.[89,90] Very recent experiments with tethered particle motion by Driessen et al. confirmed the above findings, and also showed that the increase of temperature can lead to a more compact structure of dsDNA and the temperature-dependent P is tightly coupled to the content of GC base pairs.[91] It is understandable that a dsDNA is more flexible at higher temperature because the ds helix is generally stabilized by base pairing/stacking interactions. However, a previous experiment showed that dsDNA flexibility can become weaker with increasing temperature in [2C, 20C], which was proposed to be attributed to the sequence-direct DNA curvature.[92] Such temperature-weakened flexibility for some dsDNA requires further analysis. 3.4. Sequence effect It is well known that dsDNA flexibility depends on its sequence, which directly affects dsDNA stability.[1−3, 93 , 94 ] Numerous previous studies have shown that different sequence arrangements can greatly influence the stability of dsDNA under bending and its ability to form kinks, which can induce base pair slide to form non-native contacts.[1−3] Simultaneously, the sequence can also influence the mechanical properties of a dsDNA in contact with proteins such as histones.[1−3] Olson et al. have performed the statistical analysis on X-ray crystal structures and found that different sequences produce distinct flexibility, where the AA·TT step belongs to the rigid class while GG·CC and GC·GC dinucleotides are even more flexible.[93] However, this conclusion may be invalid because of the choice of system where the central regions were rich in AT content while the GC steps were segregated at the terminals.[95] Furthermore, based on statistical analysis of X-ray crystal structures of protein-oligonucleotide complexes, Olson et al. 8 found that the average twisting of base pair steps increases in the same order within three standard chemical classes: pyrimidine–purine, purine–purine, and purine–pyrimidine.[94] Ortiz and Pablo proposed a coarse-grained model for the effect of sequence on the overall stability and flexibility of dsDNA under bending constraints.[96] They found that longer repeated segments such as AAAAAAAA were more likely to form a kink, while short repetitive segments such as CCC were less likely to form a kink. This is because a base in the AA...AA strands can slide more easily to form non-native contacts with neighboring complementary bases in the repeated TT...TT sequence.[96] In addition to the above theoretical approaches, scanning force microscopy and AFM have been widely employed to characterize the flexibility of DNA.[95,97,98] Scipioni et al. described the intrinsic curvature of DNAs based on scanning force microscopy images, confirming that A·T-rich sequences are more flexible than G·C-rich sequences.[95] To obtain a more detailed understanding of the sequence-dependent flexibility of dsDNA, Geggier and Vologodskii determined the bending persistence length more accurately based on a cyclization method of short DNA fragments, and their data showed that variations in P induced by different sequences might be sufficiently large to affect their biological activity. Such variations can also affect the binding affinity of DNA-protein complexes, in which dsDNA segment shows sharp bending.[99] Moreover, extensive theoretical and experimental analyses showed that DNA fragments containing A-tracts exhibited a tighter bend at the 3 end than at the 5 end, where the “A-tract” refers to the duplex (dA)n·(dT)n and is equivalent to the “T-tract”.[3,96,100] TA and AG·CT steps show higher roll angle values compared to GC and GG·CC steps, and can also bend DNA even in the absence of the A-tract.[100] In addition to the specific sequence described above, the sequence GGGCCC also showed a net bend.[101] With the rapid development of computational facility, molecular simulation has become an important tool for exploring the effect of sequence on the flexibility of dsDNA at the atomic level.[22,102] Based on an MD study of dsDNA in the gas phase, Xiao and Liang found that all pyrimidine rings were highly flexible in either isolated or paired states, whereas the imidazole rings were relatively more rigid.[22] Lavery et al. also performed systematic MD simulations to study the nearest-neighbor effects on base pairing.[102] Their simulations suggested that to predict the sequence dependence of DNA structure and dynamics, next-nearest-neighbor interactions should be taken into account because the effect is significant.[102] 3.5. DsDNA under high force As described above, the elastic properties of dsDNA, such as stretching modulus and 9 bending persistence length can be investigated through single-molecule force-extension experiments by fitting extension-force curves to the (extensible) WLC model.[49,54,103−109] Under low stretching force (e.g., < 20 pN), dsDNA generally maintains the B-form, and the measured properties reflect the intrinsic elasticity of B-DNA with specific sequences under specific environmental conditions. However, under high stretching force (> 65 pN, near physiological ionic conditions), apparent overstretching transitions can occur in the dsDNA structure.[104−108] Such overstretching transitions are based on the early single-molecule experiments of dsDNA and have been proposed to involve a stretching-induced melting transition from dsDNA to ssDNA.[28,49,110−112] However, very recent single-molecule experiments have revealed that much more complex overstretching transitions from dsDNA to other overstretched structures occur, such as peeled ssDNA with one strand peeled from another, DNA bubble with two strands separated internally, and S-DNA with elongated base pairs.[105−108] These experimental findings were also suggested using early theoretical models.[112] These recent experiments have yielded the following major findings:[54,104−108] 1. The overstretching transitions for dsDNA under high force strongly depend on the DNA sequence (GC-rich or AT-rich), the state of dsDNA ends (end-open or end-closed), ionic condition, and temperature.[54,104−108] 2. For end-closed dsDNA, whose ends are DNA hairpins, a high stretching force induces the transition from B-DNA to S-DNA at high ionic strength, while induces the transition from B-DNA to a DNA bubble at low ionic strength. Additionally, the overstretching force for the transitions appears to be higher for CG-rich DNA and lower for AT-rich DNA.[54,108] 3. For end-open dsDNA, whose ends are without constraints, a high force induces the overstretching transition from B-DNA to S-DNA at high ionic strength, while induces the transition from B-DNA to peeled ssDNA at low ionic strength. Additionally, the overstretching force for the transition from B-DNA to peeled ssDNA is higher for GC-rich dsDNA and lower for AT-rich dsDNA, while there is no transition between S-DNA and B-DNA for AT-rich dsDNA over the wide range of ionic strength of 1 mM to 100 mM.[108] 4. The S-DNA is a new form of dsDNA with elongated base pairs, which has been continuously stretched up to approximately 70% beyond its canonical B-form contour length and is more flexible (P ~10 nm) than B-DNA (P ~50 nm). The DNA bubble with two internal ss strands also has much higher flexibility than B-DNA.[54] Extensive single-molecule experiments have shown the overstretching transition is directly 10 coupled to the stability of a dsDNA, which strongly depends on the content of GC base pairs, ionic conditions, and end-constraints. Thus, it is expected that at lower temperature and in multivalent ion solutions, a dsDNA can be stretched to an S-DNA at lower ionic strength because of the higher stability. The atomic structures of overstretched dsDNA can be modeled by all-atom MD simulations.[24] 3.6. Flexibility of short DNA Numerous recent experiments have suggested that short DNAs have high flexibility compared with those in kilo-base pairs.[113−117] Cyclization experiments have shown that short dsDNAs of ~100 base pairs (bps) formed circles much faster than predicted by the WLC model mainly because of the unusually large local bend angle induced by kinking.[113,114] Another series of experiments using FRET and SAXS by Yuan et al. also suggested the higher flexibility of short DNAs of 15−89 bps, which were beyond the description of the conventional WLC model.[9] In addition, recent SAXS experiments of short DNAs of ≤ 35 bps with two end gold nanocrystals by Mathew-Fenn et al. suggested that short DNAs are at least one order of magnitude more extensible than long DNAs of kilo-bps revealed by previous single-molecule stretching experiments.[118] This high flexibility for short dsDNA has been proposed to be attributed to defect excitation, which may reduce the local bending energy of dsDNA through local DNA melting of a few base pairs, local DNA kinking, and excitation of a few base pairs of S-DNA.[86,112,115,119,120] Since the local stability of dsDNA depends strongly on sequence, ionic strength, and temperature,[2,30,31,63,64,93,94] such defect excitation may be sensitive to temperature and ionic strength.[112] However, a similar SAXS experiment showed that the flexibility of short DNAs of 42−94 bps with two linked gold nanocrystals could be described by the WLC model with a persistence length of ~50 nm.[11] On the illusive controversy, the atomic MDs have also been employed to probe the flexibility of short DNAs of 5−50 bps, and the results showed that shorter DNA may have higher apparent flexibility, which is attributed to the higher flexibility of ~6 bps at each end.[88] Nevertheless, how to explain the experiments with labeling nanocrystals developed by Mathew-Fenn et al. and Yuan et al. at the atomic level is still required, which would assist not only the understanding of the experimental findings, but also the understanding of the effect of labeling nanocrystals on the flexibility of short biomolecules. 4. Flexibility of RNA 4.1. Flexibility of dsRNA 11 Recently, dsRNA has been highly valued because its role in the life cycle of a cell is now more crucial than previously considered.[7,14,15,121−123] In addition to being a central role in RNA interference,[121] dsRNAs may have the potential applications in nanomedicine and nanomaterial.[122,123] Because of these important applications, the flexibility of dsRNA has been studied in various single-molecule experiments,[7,14,15] and the experimental measurements are tabulated in Table 2. Unlike dsDNA, which is generally in the B-form, dsRNA forms a thicker right-handed duplex in the “A-form”,[124] and this special helical structure gives dsRNA different flexibilities. In an early experiment of transient electric birefringence (TEB), Hagerman et al. obtained a bending persistence length of ~60 ± 10 nm for dsRNA,[125,126] which was 20−30% larger than the accepted value for dsDNA. Therefore, dsRNA is somewhat stiffer than dsDNA. Recently, Dekker et al. obtained a mean bending persistence length of 63.8 ± 0.7 nm through force-extension measurements with magnetic tweezers and of 62 ± 2 nm using AFM measurements for long dsRNAs.[7] They also obtained a torsional persistence length of ~99 ± 5 nm at an external stretching force F = 6.5 pN, which was similar to the value for dsDNA, as well as a stretching modulus of 350 ± 100 pN, which was three-fold lower than that of dsDNA. Interestingly, they observed that dsRNA exhibits positive twist-stretch coupling, which is in contrast to dsDNA with negative twist–stretch coupling.[14] Herrero-Galan et al . systematically measured the mechanical properties of dsRNA under different ion conditions at the single-molecule level.[15] They found that the values for bending persistence length P for dsRNA were consistently larger than those for dsDNA under the same ionic conditions, and that P decreased with increasing salt concentration over the range of 0−500 mM NaCl, which is similar to dsDNA. Furthermore, the OT measurements revealed that the stretching modulus S of dsRNA increased with increasing salt concentration, in agreement with the trend for dsDNA,[28] while S for dsRNA was much lower than that of dsDNA under the same salt conditions.[16,28] Therefore, compared with dsDNA, dsRNA has a larger bending persistence length and a similar torsional persistence length. However, the stretching modulus of dsRNA is nearly three-fold lower than that of dsDNA, and a surprising difference between dsRNA and dsDNA is that dsRNA has a positive twist-stretch coupling parameter while dsDNA has a negative one. The microscopic mechanism for the apparent difference in the flexibility between dsRNA and dsDNA remains unclear and requires further investigation. 4.2. Flexibility of structured RNA Generally, RNAs fold into more complex native structures rather than keep a denatured ss 12 chain or a perfect duplex.[127,128] The flexibility of complex RNA structures is important for their biological functions, such as RNA-protein recognition and gene regulation.[129,130] Understanding the flexibility of structural RNAs would enable the deep exploration of their biological functions and related applications such as structure-based drug design.[131,132] However, there have been few extensive studies on examining the flexibility of structural RNAs. In Table 3, we summarized the existing experimental measurements for the elastic properties of structured RNAs beyond the states of the ss chain and helix.[133−143] For RNA hairpins with a bulge loop (and an internal loop), Zacharias and Hagerman performed a series of TEB experiments.[134,137,138] For RNA hairpins with a bulge loop, the bending angles at the junction induced by bulge loops of different sizes and base compositions were determined, and were found to increase monotonically with the increment from ~8° to ~20° for both bulge loops of An and Un as n was increased from 1 to 6 in the absence of Mg2+. Moreover, for bulge loops constituted by Un, the presence of Mg2+ reduced the increment of bending angle by a factor of 2 for all of n values.[137] However, for RNA hairpins with symmetric internal loops with the forms An-An and Un-Un (n = 2, 4, 6), it was found that the internal loops could only distort RNAs by values that were much smaller than their bulge loop counterparts.[138] Other experimental methods have also been employed to unravel the flexibility of structured RNAs. Al-Hashimi et al. performed a series studies on the HIV-1 TAR RNA with NMR and residual dipolar coupling (RDC), in combination with coarse-grained modeling by Brooks et al.[139,144] Because of the high resolution of the method, detailed dynamic motions inside RNAs can be captured and some macroscopic quantities such as bend angle can be measured. Their experiments showed that the bending angle of HIV-1 TAR RNA decreased with increasing NaCl concentration, which was also predicted by recent coarse-grained models.[140,141,144, 145 ] Thirumalai et al. developed and employed an empirical formula for the WLC model to describe the flexibility of the Azoarcus ribozyme at different monovalent and divalent ion concentrations. Additionally, the corresponding persistence lengths were derived by fitting the empirical formula for the WLC model to the measured experimental data.[142] Rather than examining the local details of non-helical RNA elements, Fulle and Gohlke studied whole RNA molecules directly. They first modeled an RNA structure as a topological network representation using a constraint counting method, and then ran a simulation with a framework rigid optimized dynamics algorithm (FRODA). The root-mean-square fluctuations of all atoms can be determined after simulations, and then the flexibility can be estimated both locally and globally. Quantitative comparisons between FRODA simulations and NMR experiments showed good agreement for some RNAs, including tRNA and pseudoknots.[131] 13 Furthermore, the flexibility of the ribosomal exit tunnel, a large RNA-protein complex, has been studied using this method.[143] Very recently, a python-based software package named constraint network analysis (CNA) was developed by Gohlke et al.[146] Using the CNA algorithm, one can obtain both global and local properties for the input biomolecules, including RNAs. However, such constraint counting in the CNA algorithm may require more physical optimization and extensive validation. Since structured RNAs differ from DNA and RNA helices whose flexibility can be well quantified by persistence length, stretching modulus, and twisting modulus, characterizing the flexibility of structured RNAs in a straightforward and quantitative manner remains unclear. 5. Conclusion and perspective As described above, extensive experiments and theoretical modeling have revealed that the flexibility of nucleic acids is tightly coupled to several critical factors. First, the states of structures can dominate the flexibility of nucleic acids, such as the states of the ss chain, ds helices, partially melted helices, and more complicated tertiary folds, corresponding to significantly different flexibilities. Second, the sequences of nucleic acids determine their structures and stabilities, and thus strongly influence their flexibility. Third, temperature can directly determine the state of structures of nucleic acids and thus greatly influence their flexibility. Additionally, solution conditions such as metal ions, which can strongly interact with nucleic acids, significantly affect nucleic acid flexibility, particularly multivalent ions. Recent developments in computation facility and molecular force fields have enabled extensive explorations of the flexibility of nucleic acids at the atomic level. However, despite this great progress in understanding the flexibility of nucleic acids, many important elusive problems must be explored. We will discuss several major challenging issues in the following sections. Flexibility of DNA on a short length scale A recent AFM experiment showed that spontaneous large-angle bends were many times more prevalent than predicted by the WLC model,[115] suggesting that dsDNA may have much higher flexibility on the short length scale.[147] To examine the mechanism for such higher flexibility on the short length scale, atomic-level MD simulations have been employed for two short dsDNAs, and the calculated apparent persistence length can be ~20 nm on a very short scale (~1−2 bp), and exhibits the oscillation periodically between 20 nm and 100 nm.[148] A correlated WLC model has also been developed to explain the experimental findings, while the microscopic mechanism for such proposed correlation remains unclear.[ 149 ] Very recently, 14 high-resolution AFM in solution has been used to analyze the effect on a short length scale, which showed that dsDNA can be well described by the WLC model on a scale beyond 2−3 helical turns.[116,117] However, on a length scale below the threshold of 2−3 helical turns, quantification of the flexibility remains limited by the limitations of AFM. Furthermore, how the local kinking and disruption of hydrogen bond in base pairing affect the flexibility of dsDNA is still unknown. Additionally, the effect of substrate in the AFM experiments on the flexibility of dsDNA must be examined, since a theoretical modeling indicated that Mg2+-mediated attraction between DNA and substrate can cause DNA softening.[120] On a short length scale, atomistic MD can become a powerful tool, while reliable force fields in MD are essentially required. Flexibility of helices of different conformations Two typical conformations of nucleic acids include B-form DNA (B-DNA) and A-form RNA (A-RNA). DsDNA and dsRNA are generally present in B-form and A-form, respectively. B-DNA and A-RNA helices show similar helical structures and some of their elastic properties are qualitatively similar, such as bending persistence length and torsional modulus.[14,15,27] However, B-DNA and A-RNA are significantly different for some elastic properties, such as stretching modulus and stretching-twisting coupling. The stretching modulus of B-DNA is ~3 times higher than that of A-RNA.[14] More surprisingly, single-molecule stretching experiments showed that the stretching-twisting coupling parameter of B-DNA was negative, i.e., the DNA stretched by pulling force (4−8 pN) was accompanied by overwinding of the helix, while that of A-RNA was positive. However, very recently, Manning analyzed existing experimental data and concluded that DNA stretching by environmental change such as the decrease of salt concentration is companied by helix unwinding.[72] Therefore, several unanswered questions remain: (1) Why are B-DNA and A-RNA different in stretching modulus and stretching-twisting coupling? (2) How can we unify the results involving the pulling of B-DNA and the analyses based on experiments involving free B-DNA? (3) Is the flexibility of other conformations of nucleic acid helices such as A-DNA and Z-DNA, also different from the flexibility of B-DNA? Further studies, particularly those on the microscopic level, are required to answer these questions. Effect of high salt and multivalent salt Numerous experiments have shown that the persistence length of dsDNA decreases and stretching modulus increases with increasing salt concentration such as NaCl. However, the available data regarding salt-dependent P can be categorized in two ways: (1) P will not continue 15 to decrease after NaCl exceeds 0.1 M;[28,66] (2) P will continue to decrease after NaCl exceeds 0.1 M.[68,69] The classic OSF theory supports the former, while recent coarse-grained simulations and theoretical analysis based on the counterion condensation theory agree with the latter. Therefore, for the salt-dependent flexibility of DNA, several questions remain unanswered: (1) To what extent does salt in the solution influence the flexibility of DNA, and what are the relative fractions of electrostatic and intrinsic contributions to the global flexibility of DNA? (2) Why does P of DNA continue to decrease or become nearly invariant after NaCl exceeds 0.1 M? (3) Why does the stretching modulus of DNA increase at higher salt concentration? (4) Further studies are required to understand the salt dependence of other elastic properties such as torsional modulus and stretching-twisting coupling. A series of experiments in combination with molecular modeling at the atomic level is required to resolve these issues. The limited experiments on the flexibility of DNA in multivalent salt have demonstrated the dramatic roles of multivalent ions. Multivalent ions can cause an apparent decrease in the persistence length and an apparent increase in the stretching modulus of dsDNA compared with monovalent salt.[28] Such decrease in P has been attributed to multivalent ion-mediated intra-chain attractive force,[82,83,150,151] while the increase in S has not been thoroughly explained. Therefore, to systematically quantify the effect of multivalent ions such as Mg2+ and Co(NH3)6 3+ on DNA flexibility is still required, particularly on the local deformation of DNA helix induced by multivalent ions. In addition, the effect of multivalent ions on the flexibility of A-RNA has not been widely examined, while previous experiments have shown that ions can bind to an A-RNA in a very different manner to B-DNA.[152,153] This suggests that the (multivalent) ion effect on A-RNA flexibility may be significantly different from that of B-DNA and thus is highly desirable. Flexibility of RNA tertiary folds The flexibility of a DNA or RNA helix can be well described by the parameters of the elastic theory of linear polymers or elastic rods. However, RNAs are generally in the folded state of complex native structures beyond a perfect helix, and thus to quantitatively describe the flexibility of non-helix RNAs is beyond the description of the elastic parameters for a helix. The bending angle, persistence length, the distribution of radius of gyration, and the distribution of root of mean square deviation (RMSD) have been used to characterize the flexibility of non-helix RNAs. Nevertheless, there are limitations to these methods. For example, the bending angle only works well for the local bending of a helix and two-way junction. The use of persistence length for structured RNAs (e.g., tRNA) is not feasible, and may only be considered as a relative 16 quantity compared with the denatured ss state or secondary state. The distribution of the radius of gyration can be easily measured, but this value only gives a global description of the flexibility. Since a structured RNA is non-uniform in its flexibility over the entire molecule, the distribution of the radius of gyration may be inadequate for completely describing flexibility. The distribution of RMSD can describe the dynamics of all atoms and can be tracked by experiments and atomic-level modeling, while this description may be only convenient for small RNAs. A combination of these quantities may provide a thorough description of the flexibility of an RNA tertiary structure. Since RNA tertiary structures are more sensitive to temperature and ionic conditions than helices,[ 154 − 158 ] the flexibilities of RNA tertiary structures are more strongly coupled to temperature and ionic conditions, particularly multivalent ions. Furthermore, dehydrated Mg2+ and small molecules such as metabolites can interact specifically with RNAs and alter the flexibility of RNAs to aid their functions.[159−162] Therefore, understanding the flexibility of RNA tertiary structures is not well understood and requires further investigation. In summary, the results of previous studies have greatly enhanced the understanding of nucleic acid flexibility, but many questions remain and require further comprehensive investigation. In the next decade, we expect more surprising findings regarding the flexibility of nucleic acids as well as more related applications. 6. Acknowledgments We are grateful to Feng-Hua Wang, Yuan-Yan Wu and Ya-Zhou Shi for valuable discussions. This work was supported by the National Key Scientific Program (973)-Nanoscience and Nanotechnology (No. 2011CB933600), the National Science Foundation of China grants (11175132, 11575128 and 11374234), and the Program for New Century Excellent Talents (Grant No. NCET 08-0408). 17 Table 1. Experimental measurements for the flexibility of single-stranded DNAs/RNAs ss nucleic acids 13k nt ssDNA References Bosco et al[36] 10.5k nt ssDNA McIntosh et al[38] dT40, rU40 Chen et al[41] dT30 Meisburger et al[42] dT8−dT100; dA8−dA50 12−120 nt ss nucleic acidsa Wang et al[47] Sim et al[43] dT12,dT24,dA12,dA24 Mills et al[50] Poly(U) 280−5386 nt ssDNA Seol et al[51] Tinland et al[52] Ionic conditions 10−1000 mM Na+; 0.5−10 mM Mg2+ 20−3500 mM K+; 20−2000 mM Na+; 0.2−50 mM Mg2+; 0.2−50 mM Ca2+ 0−800 mM Na+; 0−100 mM Mg2+ 20 mM Na+; 0−20 mM Mg2+ 12.5−1000 mM Na+ 1−1000 mM Na+; 0.03−300 mM Mg2+; 0.01−100 mM Co3+ 8−64 mM Na+; 0−8 mM Mg2+ 5−500 mM Na+ 1−100 mM EDTA Thermodynamic quantities S, L, P P Ree, L, P Rg, Ree Rg, P Ree, P P S, P Rg, P a This was a computational study that collected various experimental data for P of ss nucleic acids; Rg: radius of gyration; Ree: end-to-end distance; L: contour length; P: bending persistence length; S: stretching modulus. 18 Table 2. Experimental measurements for the flexibility of dsDNAs/dsRNAs dsDNAs or dsRNAs References Ionic conditions & Thermodynamic 4.2k bp & 8.3k bp dsRNA Abels et al[7] 16, 21, 66 & 89 bp dsDNA Yuan et al[9] 10, 15, 20, 25, 30, 35 bp Mathew-Fenn et al[10] dsDNA Temperature moderate salt buffer 500 mM Na+ 100 mM Na+ quantities P Rg, P Ree, 4.2k bp dsRNA & 3.4k bp Lipfert et al[14] 100 mM & 320 mM Na+ P, C, S, D Herrero-Galan et al[15] 0−500 mM Na+ Baumann et al[28] dsDNA 4k bp λDNA & 4k bp dsRNA 50k bp λDNA 200 bp λDNA Geggier et al[31] 14.8k bp dsDNA 50k bp λDNA T7 DNA Bryant et al[58] Strick et al[62] Sobel et al[68] Borochov[69] 6954±20 bp dsDNA dsDNA (125bp−23000 bp) Mantelli et al[73] 3888 bp dsDNA pBR322 dsDNA & Φ6 dsRNA 685 bp dsDNA Wang et al[74] Lang et al[76] Driessen et al[91] blunt-ended DNA Porschke[92] fragments (41−256 bp) P, S P, S P C P, C Ree, P Rg, P P P, S, L L P P, RH 1.86−586 mM Na+; Mg2+,Put2+,Spd3+ & Co(NH3)6 3+ TBE buffer; 5−60℃ 100 mM Na+ 10 mM phosphate buffer 5−3000 mM Na+ 7.3−4000 mM Na+ 1 mM Mg2+ & 1−100 mM Na+ Na+, K+, Mg2+, Spd3+ 48−500 mM NH4Cl 60 mM K+; 100 & 150 mM Na+; 23−52°C 2.4−110 mM Na+; 0.1 & 10 mM Mg2+; 2−20℃ 200 bp dsDNA 2743 bp dsDNA Geggier et al[99] Wiggins et al[115] TBE buffer 12 mM Mg2+ P Bend angle S: stretching modulus; P: bending persistence length; C: torsional persistence length; L: contour length; Ree: end-to-end distance; Rg: radius of gyration; D: twist–stretch coupling parameter; : variance of Ree; RH: hydrodynamic radius. 19 2R2R Table 3. Different methods employed to probe the flexibilities of structural RNAs RNAs tRNA (Asp) References Fulle et al[131] tRNA (Phe) Roh et al[133] Methods Thermodynamic quantities Constraint counting network & FRODA simulation Quasielastic neutron scattering spectroscopy RMSD Rg, P HIV-1 TAR RNA subsequence of a sRNA (DsrA) Zacharias et al[134] Gel electrophoresis & TEB Bend angle De Almeida Ribeiro E et al[135] SAXS & NMR & ensemble optimization Rg bacterial ribosomal Fulle et al[136] Normal mode analysis & Binding free energies, method A-site RNA MD simulation segments of bulge Zacharias et al[137] TEB loops segments of symmetric internal loops HIV-1 TAR RNA HIV-1 TAR RNA Azoarcus ribozyme & RNase P Zacharias et al[138] TEB Al-Hashimi et al[139] NMR & RDC & MD Al-Hashimi et al[140][141] Caliskan et al[142] NMR & RDC SAXS & WLC ribosomal exit tunnel Fulle et al[143] Constraint counting network & FRODA simulation RMSD Bend angle Bend angle RMSD RMSD P RMSD P: bending persistence length; Rg: radius of gyration; RMSD: Root-mean-square deviation. 20 References [1] Hagerman P J 1988 Annu. Rev. Biophys. Biophys. Chem. 17(1) 265 [2] Peters J P and Maher L J 2010 Q. Rev. Biophys. 43(01) 23 [3] Hagerman P J 1990 Annu. Rev. Biochem. 59(1) 755 [4] Tan Z J and Chen S J 2011 Met Ions Life Sci 9 101 [5] Lebrun A, Shakked Z and Lavery R 1997 Proc. Natl. Acad. Sci. USA 94 2993 [6] Mazur A K and Maaloum M 2014 Phys. Rev. Lett. 112(6) 068104 [7] Abels J A, Moreno-Herrero F, Van der Heijden T, Dekker C and Dekker N H 2005 Biophys. J. 88(4) 2737 [8] Kang J, Jung J and Kim S K 2014 Biophys. Chem. 195 49 [9] Yuan C, Chen H, Lou X W and Archer L A 2008 Phys. Rev. Lett. 100(1) 018102 [10] Mathew-Fenn R S, Das R and Harbury P A 2008 Science 322(5900) 446 [11] Mastroianni A J, Sivak D A, Geissler P L and Alivisatos A P 2009 Biophys. J. 97(5) 1408 [12] Wang X L, Zhang X H, Cao M, Zheng H Z, Xiao B, Wang Y and Li M 2009 J. Phys. Chem. B 113(8) 2328 [13] Li W, Wang P Y, Yan J and Li M 2012 Phys. Rev. Lett. 109(21) 218102 [14] Lipfert J, Skinner G M, Keegstra J M, Hensgens T, Jager T, Dulin D, Köbera M, Yu Z B, Donkers S P, Chou F C, Das R and Dekker N H 2014 Proc. Natl. Acad. Sci. USA 111(43) 15408 [15] Herrero-Galan E, Fuentes-Perez M E, Carrasco C, Valpuesta J M, Carrascosa J L, Moreno-Herrero F and Arias-Gonzalez J R 2013 J. Am. Chem. Soc. 135 122 [16] Lipfert J, Kerssemakers J W, Jager T and Dekker N H 2010 Nat. Methods 7(12) 977 [17] Marko J F and Siggia E D 1995 Macromolecules 28(26) 8759 [18] Toan N M and Thirumalai D 2012 J. Chem. Phys. 136(23) 235103 [19] Song T J and Liang H J 2012 J. Am. Chem. Soc. 134(26) 10803 [20] Zhang Y J, Zhang J and Wang W 2011 J. Am. Chem. Soc. 133(18) 6882 [21] Xiao S Y, Zhu H, Wang L and Liang H J 2014 Soft matter 10(7) 1045 [22] Xiao S Y and Liang H J 2012 J. Chem. Phys. 136(20) 205102 [23] Zhao Y J, Huang Y Y, Gong Z, Wang Y J, Man J F and Xiao Y 2012 Sci. Rep. 2 734 [24] Qi W P, Lei X L and Fang H P 2010 ChemPhysChem 11(10) 2146 [25] Tan Z J and Chen S J 2008 Biophys. J. 94 3137 [26] Hou X M, Zhang X H, Wei K J, Ji C, Dou S X, Wang W C, Li M and Wang P Y 2009 Nucleic Acids Res. 37(5) 1400 [27] Wenner J R, Williams M C, Rouzina I and Bloomfield V A 2002 Biophys. J. 82(6) 3160 [28] Baumann C G, Smith S B, Bloomfield V A and Bustamante C 1997 Proc. Natl. Acad. Sci. USA 94(12) 6185 [29] Gray H B and Hearst J E 1968 J. Mol. Biol. 35(1) 111 [30] Forties R A, Bundschuh R and Poirier M G 2009 Nucleic Acids Res. 37(14) 4580 [31] Geggier S, Kotlyar A and Vologodskii A 2011 Nucleic Acids Res. 39(4) 1419 [32] Hannon G J 2002 Nature 418(6894) 244 [33] Saenger W 1984 Principles of nucleic acid structure (New York: Springer-Verlag) [34] Diebold S S, Kaisho T, Hemmi H, Akira S and e Sousa C R 2004 Science 303(5663) 1529 [35] Sung P and Robberson D L 1995 Cell 82(3) 453 [36] Bosco A, Camunas-Soler J and Ritort F 2013 Nucleic Acids Res. 42(3) 2064 [37] Landy J, McIntosh D B and Saleh O A 2012 Phys. Rev. Lett. 109(4) 048301 [38] McIntosh D B and Saleh O A 2011 Macromolecules 44(7) 2328 [39] Seol Y, Skinner G M, Visscher K, Buhot A and Halperin A 2007 Phys. Rev. Lett. 98(15) 158103 [40] Zhang Y, Zhou H J and Ou-Yang Z C 2001 Biophys. J. 81(2) 1133 [41] Chen H, Meisburger S P, Pabit S A, Sutton J L, Webb W W and Pollack L 2012 Proc. Natl. Acad. Sci. USA 109(3) 799 [42] Meisburger S P, Sutton J L, Chen H, Pabit S A, Kirmizialtin S, Elber R and Pollack L 2013 Biopolymers 99(12) 1032 [43] Sim A Y, Lipfert J, Herschlag D and Doniach S 2012 Phys. Rev. E 86(2) 021901 [44] Ke C, Humeniuk M, Hanna S and Marszalek P E 2007 Phys. Rev. Lett. 99(1) 018302 [45] Toan N M and Thirumalai D 2012 J. Chem. Phys. 136(23) 235103 [46] Sing C E and Alexander-Katz A 2011 Macromolecules 44(17) 6962 [47] Wang F H, Wu Y Y and Tan Z J 2013 Biopolymers 99(6) 370 [48] Dessinges M N, Maier B, Zhang Y, Peliti M, Bensimon D and Croquette V 2002 Phys. Rev. Lett. 89(24) 248102 [49] Smith S B, Cui Y and Bustamante C 1996 Science 271(5250) 795 21 [50] Mills J B, Vacano E and Hagerman P J 1999 J. Mol. Biol. 285(1) 245 [51] Seol Y, Skinner G M and Visscher K 2004 Phys. Rev. Lett. 93(11) 118102 [52] Tinland B, Pluen A, Sturm J and Weill G 1997 Macromolecules 30(19) 5763 [53] Tan Z J and Chen S J 2008 Biophys. J. 95(2) 738 [54] Cocco S, Yan J, Léger J F, Chatenay D and Marko J F 2004 Phys. Rev. E 70(1) 011910 [55] Sim A Y, Lipfert J, Herschlag D and Doniach S 2012 Phys. Rev. E 86(2) 021901 [56] Riemer S C and Bloomfield V A 1978 Biopolymers 17(3) 785 [57] T J Richmond, J T Finch, B Rushton, D Rhodes and A Klug 1984 Nature 311 532 [58] Bryant Z, Stone M D, Gore J, Smith S B, Cozzarelli N R and Bustamante C 2003 Nature 424(6946) 338 [59] Horowitz D S and Wang J C 1984 J. Mol. Biol. 173(1) 75 [60] Shore D and Baldwin R L 1983 J. Mol. Biol. 170(4) 983 [61] Vologodskii A V and Marko J F 1997 Biophys. J. 73(1) 123 [62] Strick T R, Bensimon D and Croquette V 1999 Genetica 106(1-2) 57 [63] Tan Z J and Chen S J 2006 Biophys. J. 90(4) 1175 [64] Tan Z J and Chen S J 2007 Biophys. J. 92(10) 3615 [65] Peters J P, Mogil L S, McCauley M J, Williams M C and Maher L J 2014 Biophys. J. 107(2) 448 [66] Maret G and Weill G 1983 Biopolymers 22(12) 2727 [67] Podesta A, Indrieri M, Brogioli D, Manning G S, Milani P, Guerra R, Finzi L and Dunlap D 2005 Biophys. J. 89(4) 2558 [68] Sobel E S and Harpst J A 1991 Biopolymers 31(13) 1559 [69] Borochov N, Eisenberg H and Kam Z 1981 Biopolymers 20(1) 231 [70] Tan Z J and Chen S J 2005 J. Chem. Phys. 122(4) 044903 [71] Tan Z J and Chen S J 2009 Methods Enzymol. 469 465 [72] Manning G S 2015 Biopolymers 103(4) 223 [73] Mantelli S, Muller P, Harlepp S and Maaloum M 2011 Soft Matter 7(7) 3412 [74] Wang M D, Yin H, Landick R, Gelles J and Block S M 1997 Biophys. J. 72(3) 1335 [75] Anderson P and Bauer W 1978 Biochemistry 17(4) 594 [76] Lang D, Steely H T, Kao C Y and Ktistakis N T 1987 Biochim. Biophys. Acta. 910(3) 271 [77] Odijk T 1977 J. Polym. Sci. Polym. Phys. Ed. 15(3) 477 [78] Skolnick J and Fixman M 1977 Macromolecules 10(5) 944 [79] Manning G S 2006 Biophys. J. 91(10) 3607 [80] Savelyev A, Materese C K and Papoian G A 2011 J. Am. Chem. Soc. 133(48) 19290 [81] Savelyev A 2012 Phys. Chem. Chem. Phys. 14(7) 2250 [82] Tan Z J and Chen S J 2006 Nucleic Acids Res. 34(22) 6629 [83] Tan Z J and Chen S J 2012 Biophys. J. 103(4) 827 [84] Rau D C and Parsegian V A 1992 Biophys. J. 61(1) 246 [85] Schellman J A 1974 Biopolymers 13(1) 217 [86] Chen H and Yan J 2008 Phys. Rev. E 77(4) 041907 [87] Chen H, Liu Y, Zhou Z, Hu L, Ou-Yang Z C and Yan J 2009 Phys. Rev. E 79(4) 041926 [88] Wu Y Y, Bao L, Zhang X and Tan Z J 2015 J. Chem. Phys. 142(12) 125103 [89] Kochoyan M, Leroy J L and Guéron M 1987 J. Mol. Biol. 196(3) 599 [90] Coman D and Russu I M 2005 Biophys. J. 89(5) 3285 [91] Driessen R P, Sitters G, Laurens N, Moolenaar G F, Wuite G J, Goosen N and Dame R T 2014 Biochemistry 53(41) 6430 [92] Porschke D 1991 Biophys. Chem. 40(2) 169 [93] Olson W K, Marky N L, Jernigan R L and Zhurkin V B 1993 J. Mol. Biol. 232(2) 530 [94] Olson W K, Gorin A A, Lu X J, Hock L M and Zhurkin V B 1998 Proc. Natl. Acad. Sci. USA 95(19) 11163 [95] Scipioni A, Anselmi C, Zuccheri G, Samori B and De Santis P 2002 Biophys. J. 83(5) 2408 [96] Ortiz V and de Pablo J J 2011 Phys. Rev. Lett. 106(23) 238107 [97] Rivetti C, Guthold M and Bustamante C 1996 J. Mol. Biol. 264(5) 919 [98] Le Cam E, Culard F, Larquet E, Delain E and Cognet J A 1999 J. Mol. Biol. 285(3) 1011 [99] Geggier S and Vologodskii A 2010 Proc. Natl. Acad. Sci. USA 107(35) 15421 [100] MacDonald D, Herbert K, Zhang X, Polgruto T and Lu P 2001 J. Mol. Biol. 306(5) 1081 [101] Brukner I, Dlakic M, Savic A, Susic S, Pongor S and Suck D 1993 Nucleic Acids Res. 21(4) 1025 [102] Lavery R, Zakrzewska K, Beveridge D, Bishop T C, Case D A, Cheatham III T, Dixit S, Jayaram B, Lankas F, Laughton C, Maddocks J H, Michon A, Osman R, Orozco M, Perez A, Singh T, Spackova N and 22 Sponer J 2010 Nucleic Acids Res. 38(1) 299 [103] Cluzel P, Lebrun A, Heller C, Lavery R, Viovy J L, Chatenay D and Caron F 1996 Science 271(5250) 792 [104] Bustamante C, Bryant Z and Smith S B 2003 Nature 421(6921) 423 [105] Fu H, Chen H, Zhang X, Qu Y, Marko J F and Yan J 2011 Nucleic Acids Res. 39(8) 3473 [106] Zhang X, Chen H, Fu H, Doyle P S and Yan J 2012 Proc. Natl. Acad. Sci. USA 109(21) 8103 [107] Zhang X, Chen H, Le S, Rouzina I, Doyle P S and Yan J 2013 Proc. Natl. Acad. Sci. USA 110(10) 3865 [108] Zhang X, Qu Y, Chen H, Rouzina I, Zhang S, Doyle P S and Yan J 2014 J. Am. Chem. Soc. 136(45) 16073 [109] Li J, Wijeratne S S, Qiu X and Kiang C H 2015 Nanomaterials 5(1) 246 [110] Rouzina I and Bloomfield V A 2001 Biophys. J. 80(2) 882 [111] Rouzina I and Bloomfield V A 2001 Biophys. J. 80(2) 894 [112] Yan J and Marko J F 2004 Phys. Rev. Lett. 93(10) 108108 [113] Cloutier T E and Widom J 2004 Mol. Cell. 14(3) 355 [114] Vafabakhsh R and Ha T 2012 Science 337(6098) 1097 [115] Wiggins P A, Van Der Heijden T, Moreno-Herrero F, Spakowitz A, Phillips R, Widom J, Dekker C and Nelson P C 2006 Nat. Nanotech. 1(2) 137 [116] Mazur A K and Maaloum M 2014 Phys. Rev. Lett. 112(6) 068104 [117] Mazur A K and Maaloum M 2014 Nucleic Acids Res. 42(22) 14006 [118] Mathew-Fenn R S, Das R and Harbury P A 2008 Science 322(5900) 446 [119] Yan J, Kawamura R and Marko J F 2005 Phys. Rev. E 71(6) 061905 [120] Padinhateeri R and Menon G I 2013 Biophys. J. 104(2) 463 [121] Timmons L and Fire A 1998 Nature 395(6705) 854 [122] Guo P 2010 Nat. Nanotech. 5(12) 833 [123] Bumcrot D, Manoharan M, Koteliansky V and Sah D W 2006 Nat.Chem. Biol. 2(12) 711 [124] Metzler D 2003 Biochemistry: the chemical reactions of living cells 2nd edn. (Amsterdam: Elsevier) [125] Gast F U and Hagerman P J 1991 Biochemistry 30 4268 [126] Kebbekus P, Draper D E and Hagerman P 1995 Biochemistry 34(13) 4354 [127] Shi Y Z, Wu Y Y, Wang F H and Tan Z J 2014 Chin. Phys. B 23(7) 078701 [128] Tan Z J, Zhang W B, Shi Y Z, Wang F H 2015 Adv. Expt. Med. Biol. 827 143 [129] Frank J and Agrawal R K 2000 Nature 406(6793) 318 [130] Mandal M and Breaker R R 2004 Nat. Rev. Mol. Cell Biol. 5(6) 451 [131] Fulle S and Gohlke H 2008 Biophys. J. 94(11) 4202 [132] Murchie A I, Davis B, Isel C, Afshar M, Drysdale M J, Bower J, Potter A J, Starkey I D, Swarbrick T M, Mirza S, Prescott C D, Vaglio P, Aboul-ela F and Karn J 2004 J. Mol. Biol. 336(3) 625 [133] Roh J H, Tyagi M, Briber R M, Woodson S A and Sokolov A P 2011 J. Am. Chem. Soc. 133(41) 16406 [134] Zacharias M and Hagerman P J 1995 Proc. Natl. Acad. Sci. USA 92(13) 6052 [135] De Almeida Ribeiro E, Beich-Frandsen M, Konarev P V, Shang W F, Večerek B, Kontaxis G, Hämmerle H, Peterlik H, Svergun D I, Bläsi U and Djinović-Carugo K 2012 Nucleic Acids Res. 40(16) 8072 [136] Fulle S and Gohlke H 2010 J. Mol. Recognit. 23(2) 220 [137] Zacharias M and Hagerman P J 1995 J. Mol. Biol. 247(3) 486 [138] Zacharias M and Hagerman P J 1996 J. Mol. Biol. 257(2) 276 [139] Salmon L, Bascom G, Andricioaei I and Al-Hashimi H M 2013 J. Am. Chem. Soc. 135(14) 5457 [140] Al-Hashimi H M, Gosser Y, Gorin A, Hu W, Majumdar A and Patel D J 2002 J. Mol. Biol. 315(2) 95 [141] Zhang Q, Stelzer A C, Fisher C K and Al-Hashimi H M 2007 Nature 450(7173) 1263 [142] Caliskan G, Hyeon C, Perez-Salas U, Briber R M, Woodson S A and Thirumalai D 2005 Phys. Rev. Lett. 95(26) 268303 [143] Fulle S and Gohlke H 2009 Methods 49(2) 181 [144] Mustoe A M, Brooks C L and Al-Hashimi H M 2014 Annu. Rev. Biochem. 83 441 [145] Shi Y Z, Wang F H, Wu Y Y and Tan Z J 2014 J. Chem. Phys. 141(10) 105102 [146] Pfleger C, Rathi P C, Klein D L, Radestock S and Gohlke H 2013 J. Chem. Inf. Model. 53(4) 1007 [147] Wiggins P A, Phillips R and Nelson P C 2005 Phys. Rev. E 71(2) 021909 [148] Noy A and Golestanian R 2012 Phys. Rev. Lett. 109(22) 228101 [149] Xu X, Thio B J R and Cao J 2014 J. Phys. Chem. Lett. 5(16) 2868 [150] Widom J and Baldwin R L 1980 J. Mol. Biol. 144(4) 431 [151] Bloomfield V A 1997 Biopolymers 44(3) 269 [152] Tolokh I S, Pabit S A, Katz A M, Chen Y, Drozdetski A, Baker N, Pollack L and Onufriev A V 2014 Nucleic Acids Res. 42(16) 10823 23 [153] Wu Y Y, Zhang Z L, Zhang J S, Zhu X L, Tan Z J 2015 Nucleic Acids Res. 43(12) 6156 [154] Lipfert J, Doniach S, Das R and Herschlag D 2014 Annu. Rev. Biochem. 83 813 [155] Draper D E, Grilley D and Soto A M 2005 Annu. Rev. Biophys. Biomol. Struct. 34 221 [156] Woodson S A 2005 Curr. Opin. Chem. Biol. 9(2) 104 [157] Tan Z J and Chen S J 2010 Biophys. J. 99(5) 1565 [158] Tan Z J and Chen S J 2011 Biophys. J. 101(1) 176 [159] Serganov A and Nudler E 2013 Cell 152(1) 17 [160] Cech T R and Steitz J A 2014 Cell 157(1) 77 [161] Chen J W and Zhang W B 2012 J. Chem. Phys. 137(22) 225102 [162] Gong S, Wang Y J and Zhang W B 2015 J. Chem. Phys. 142(1) 015103 24
1307.6183
1
1307
2013-07-23T18:23:22
Slow sedimentation and deformability of charged lipid vesicles
[ "physics.bio-ph", "cond-mat.soft", "physics.comp-ph" ]
The study of vesicles in suspension is important to understand the complicated dynamics exhibited by cells in vivo and in vitro. We developed a computer simulation based on the boundary-integral method to model the three dimensional gravity-driven sedimentation of charged vesicles towards a flat surface. The membrane mechanical behavior was modeled using the Helfrich Hamiltonian and near incompressibility of the membrane was enforced via a model which accounts for the thermal fluctuations of the membrane. The simulations were verified and compared to experimental data obtained using suspended vesicles labelled with a fluorescent probe, which allows visualization using fluorescence microscopy and confers the membrane with a negative surface charge. The electrostatic interaction between the vesicle and the surface was modeled using the linear Derjaguin approximation for a low ionic concentration solution. The sedimentation rate as a function of the distance of the vesicle to the surface was determined both experimentally and from the computer simulations. The gap between the vesicle and the surface, as well as the shape of the vesicle at equilibrium were also studied. It was determined that inclusion of the electrostatic interaction is fundamental to accurately predict the sedimentation rate as the vesicle approaches the surface and the size of the gap at equilibrium, we also observed that the presence of charge in the membrane increases its rigidity.
physics.bio-ph
physics
Slow sedimentation and deformability of charged lipid vesicles Iv´an Rey Su´arez1, Chad Leidy2, Gabriel T´ellez2, Guillaume Gay3, Andres Gonzalez-Mancera1,∗ 1 Department of Mechanical Engineering, Universidad de los Andes, Bogota, Colombia. 2 Department of Physics, Universidad de los Andes, Bogota, Colombia. 3 Universite Paul Sabatier Toulose III, Toulose - France ∗ E-mail: Corresponding [email protected] Abstract The study of vesicles in suspension is important to understand the complicated dynamics exhibited by cells in in vivo and in vitro. We developed a computer simulation based on the boundary-integral method to model the three dimensional gravity-driven sedimentation of charged vesicles towards a flat surface. The membrane mechanical behavior was modeled using the Helfrich Hamiltonian and near incompressibility of the membrane was enforced via a model which accounts for the thermal fluctuations of the membrane. The simulations were verified and compared to experimental data obtained using suspended vesicles labelled with a fluorescent probe, which allows visualization using fluorescence microscopy and confers the membrane with a negative surface charge. The electrostatic interaction between the vesicle and the surface was modeled using the linear Derjaguin approximation for a low ionic concentration solution. The sedimentation rate as a function of the distance of the vesicle to the surface was determined both experimentally and from the computer simulations. The gap between the vesicle and the surface, as well as the shape of the vesicle at equilibrium were also studied. It was determined that inclusion of the electrostatic interaction is fundamental to accurately predict the sedimentation rate as the vesicle approaches the surface and the size of the gap at equilibrium, we also observed that the presence of charge in the membrane increases its rigidity. Introduction Lipid vesicles in the same size range as cells are widely used to investigate the physical properties of cell membranes, providing a simplified model system to investigate the role that the membrane plays in the mechanical deformation of cells. It is important to consider lipid membrane mechanics in order 3 1 0 2 l u J 3 2 ] h p - o i b . s c i s y h p [ 1 v 3 8 1 6 . 7 0 3 1 : v i X r a 2 to understand many relevant processes involving cell deformation. For example, red blood cells (RBCs) undergo large deformations in the microcirculation [1] which appears to optimize gas exchange between blood and the surrounding tissues [2, 3]. Given that unilamellar lipid vesicles share some features with RBCs, such as the ability to generate equilibrium biconcave shapes and dynamics under shear flow such as tank-threading and tumbling, the study of its behavior under flow conditions can provide useful information in the understanding of blood rheology [4–6]. Given the high stretching elastic modulus of lipid membranes, these are generally assumed to be incompressible (constant area per lipid molecule) [7–9]. As a consequence, the dynamics of vesicles under √ flow are sensitive to the value of the reduced volume, τ = 6 πV /S3/2 (V is the volume of the vesicle and S its surface area) [10], which quantifies the amount of area available for the deformation of the vesicle. The maximum value τ = 1 corresponds to a sphere and values of τ < 1 correspond to shapes with an initial excess of area (as compared to a sphere). In order to enforce local incompressibility a local isotropic tension, in the form of a Lagrange multiplier, can be added to the force density acting on the membrane [11–16]. The value of the Lagrange multiplier can be calculated approximately by considering a nearly incompressible elastic membrane and calculating the tension due to deformation, or by enforcing that the resulting flow field has zero divergence on the surface of the membrane. Either method neglects the effect of thermal fluctuations which is generally justified due to the relatively high stresses the hydrodynamics impose on the deforming vesicles. Lipid membrane deformability influences the movement of vesicles close to a surface [17]. For example, it was shown that deflated vesicles under shear flow at a critical shear rate are lifted from the surface by means of a force of viscous origin. After lifting from the surface, the vesicles acquire the shape of prolate ellipsoids as they move in the direction of the flow. A direct numerical simulation (DNS) using a spectral boundary integral equation method was used to quantitatively determine the lift force acting on deflated vesicles near a wall [16]. It was demonstrated that lift velocity strongly depends on the vesicle's reduced volume and the viscosity ratio. Despite the fact that the dynamic behavior of vesicles under external flow fields has been widely studied experimentally [18–20], computationally [13, 16, 21], and theoretically [12, 22], sedimentation of lipid vesicles had received little attention. Recently, Boedec and coworkers studied the sedimentation of vesicles using computational [15] and theoretical [23] methods. In both cases the contribution of thermal fluctuations is neglected and sedimentation is studied in an infinite space. Experimental observations of 3 pear-like shapes and microtether extrusion of sedimenting vesicles have also been reported [24]. Buoyancy-driven motion of drops has been widely studied in the fluid mechanics literature both ex- perimentally and numerically. For example, the buoyancy-driven interaction of viscous drops was studied experimentally and numerically using the boundary integral method [25]. The effect of surfactants on buoyancy-driven motion and interaction of viscous deformable drops was also studied using the boundary- integral method [26–29]. In all these works, the computational method has proven to be precise and in good agreement with both theoretical predictions and experimental data. In this work we focus on the gravity induced sedimentation and deformation of initially quasi-spherical vesicles as they approach a flat horizontal no-slip surface. For this purpose we develop a computational algorithm based on the boundary integral method [30]. Our algorithm is mostly based on the method described by Zinchenko and coworkers [31,32] which was developed to study the interaction of deformable drops. More recently a similar algorithm has been used to study the dynamics of a vesicle in shear flow [10] and in a wall-bound shear flow [16]. The assumption of membrane incompressibility is based on the fact that direct area expansion is expensive from an energy point of view. Studies carried out by micropipette aspiration have shown that vesicles can undergo strain at low tensions. For example 1-stearoyl-2-oleoyl-phosphatidylcholine (SOPC) vesicles have shown strains of up to 0.06 for applied tensions of 0.5 mN/m [33], this strain is accounted for by considering the smoothing of suboptical thermal fluctuations of the membrane in the low tension regime [34]. Therefore, even though changes in the lipid surface density are not expected to be significant, small membrane strain is likely as a result of smoothing of thermal undulations, which are accessible under low tension conditions. In other words, the thermal undulations act as an area reservoir which allows the membrane to deform under moderate forces. In our model, membrane mechanics are dictated by the traditional Helfrich Hamiltonian [35] and we take into account the smoothing of thermal undulations by relating area strain to tension by the constitutive function proposed by Evans and Rawicz [34]. This equation considers area strain to be caused by the superposition of the smoothing of thermal undulations plus a direct expansion of the area per molecule. The latter requires much higher energy and is not available at low tensions, such as those exerted by gravity in our experiments. Many experiments with lipid vesicles are performed in glucose/sucrose solutions on glass substrates, using fluorescent dyes to label the membrane. These studies do not control for ionic conditions, and usu- 4 ally ignore the influence of electrostatic interactions between different components in their experimental setups. These electrostatic interactions may affect the behavior of a suspended vesicle, for instance in determining the sedimentation rate close to the wall and the gap between the vesicle and the glass sur- face at equilibrium. In our work electrostatic interaction between the membrane and the planar surface are taken into account by using the Derjaguin, Landay, Verwey, Overbeek (DLVO) theory of colloidal stability [36–38], where the linear Derjaguin solution for the Poisson-Boltzmann equation of the system was used to find the interaction force between the vesicle and the substrate. By using single plane imaging microscopy (SPIM) we study the sedimentation dynamics of vesicles, this technique allows to follow a vesicle as a function of time with a better time resolution than a scanning technique such as confocal microscopy. We study the shapes and the size of the gap between the vesicle and the surface at equilibrium using confocal microscopy. The objective of this work is to develop a robust computational model, which is able to accurately predict the behavior of charged vesicles sedimenting towards a charged flat surface. The model should contain the basic physics in order to be able to predict variables such as sedimentation rate as the vesicle approaches the surface and the gap between the vesicle and the glass surface at equilibrium. Materials and methods Vesicle preparation and experimental setup 1-Palmitoyl 2-oleoyl-phosphatidylcholine (POPC) was purchased from Avanti Polar Lipids (Alabaster, AL). N -(7-nitrobenz-2-oxa-1, 3 diazol-4-yl)-1, 2 dihexadecanoyl-sn-glycero-3-phosphoethanolamine, ti- ethylammonium salt (NBD-PE) was purchased from Molecular Probes (Invitrogen, Copenhagen, Den- mark). The sucrose and glucose as well as the calcein were purchased from Sigma (Saint Louis, MS). All the aqueous solutions were prepared in milliQ water. Giant unilamellar vesicles were prepared via electroformation as described in the literature [39]. First, platinum wire electrodes inserted into Teflon wells are painted with POPC lipids dissolved in chloroform at 2 mg/mL (2-4 µL) and labeled with 1 mol% NBD-PE. The electrodes are then allowed to dry in vacuum overnight. 1 mL of sucrose solution (1 Molar) is then added to the wells and the electrodes are connected to an AC signal generator producing a sinusoidal 2V peak to peak signal at 10 Hz for 1.5 hrs. 100 µL of solution containing vesicles is taken from the well, and the vesicles are then resuspended in 300 5 µL of a glucose solution with the same osmolarity as the sucrose solution (1 Molar). This sucrose-glucose system results in the precipitation of the vesicles due to density differences between the glucose and sucrose solutions. The sedimentation of vesicles was studied through a homebuilt SPIM (UMR5088 Universite Paul Sabatier, Toulouse III and CNRS) [40]. First the glucose solution is placed into a quartz cuvette, fol- lowed by careful injection of the sucrose solution containing the vesicles. Once the sample is prepared, precipitating vesicles are localized with a 20X objective in air and a series of images are acquired through a CCD camera with a 15 seconds interval between each image. Confocal microscopy and image analysis Vesicles at equilibrium were studied using confocal microscopy using an Olympus FV1000 (Tokio, Japan). Calcein was added to the glucose solution at a 1 µM concentration in order to improve contrast. The calcein provides the external solution with counterions at an approximately 2 µM concentration. In order to measure the deformation of the vesicles, the confocal stacks were analyzed using an in-house algorithm by the following method: in each image plane, a threshold was applied, and the boundaries of the vesicle were detected using the function cvFindContours of the OpenCV library [41, 42]. The cvFitEllipse2 function from the same library was then used on the detected contours to retrieve the dimensions and positions of the vesicles found in the given plane. A KMean algorithm was applied on the vesicle centers to track each vesicle from plane to plane throughout the stack. After this automated detection, the contour of the vesicle along the z axis is defined by its borders in each image plane. The Nelder-Mead simplex algorithm [42] is then used to fit the z contour of the vesicle with a function given by: r = 1 + sin2 θ 2 1 + cos2 θ 2 (cid:18) (cid:19)α(cid:18) (cid:19)β (1) The center of the polar coordinates in the z axis corresponds to the plane where the vesicle width is maximum. Best fit is obtained by finding the best values of parameters α and β and minimizing the least-square error. From this fit, area strain and curvature are determined. To measure the equilibrium distance we waited for 15 minutes after putting the sample to start taking images of the vesicles. After waiting for this period of time no significant movement of the vesicles was observed. The whole experiment typically lasted for an hour. We performed these experiments for unlabeled POPC vesicles and for NBD-PE labeled POPC vesicles in solutions with no added salt and in 6 solutions with 3 mM NaCl concentration. Fluid mechanics model The system considered in this paper is a giant lipid vesicle (diameter ≈ 20µm) immersed into a semi- infinite external fluid. The vesicle is settling due to a gravitational force towards a no-slip, infinite planar surface. A schematic of the system is shown in Figure 1. Fluid 1, which is internal to the vesicle, and fluid 2, in which the vesicle is suspended are both assumed to be Newtonian. The fluids are separated by a phospholipid bilayer membrane which introduces a jump in the stress field between both fluids. Gravitational acceleration g is assumed to act along the z-axis, normal to the planar surface. R0, is the radius of the quasi-spherical undeformed vesicle. ρi (i = 1, 2) is the density of each fluid. The viscosities of fluids 1 and 2 are λµ and µ, respectively, where λ is the viscosity contrast between the internal and external fluids. All equations are presented in dimensionless form unless otherwise noted. Distances are scaled using R0. The characteristic time and traction are, respectively, t0 = µ(1 + λ) ∆ρgR0 , and f0 = κb R3 0 (2) where ∆ρ = ρ1 − ρ2 and κb is the bending rigidity modulus of the membrane. The analysis presented in this work is based upon the creeping motion approximation in which the inertial terms in the equation of motion are neglected entirely for vanishing small Reynolds numbers. In this limit the Stokes equations can be used to model the flow of fluids 1 and 2. During sedimentation the membrane of the vesicle is a moving interface, whose position is unknown a priori. The fluid–structure interaction problem, the Stokes equations coupled to the membrane mechanics equations, can be solved using the boundary integral formalism [31, 43] which yields in non-dimensional form (cid:90) uj = 2 g0 (cid:90) ∆fiGijdS + 2κ uiTijknkdS (3) s s where u is the velocity at each node over the membrane, κ = (λ − 1)/(λ + 1) and g0 = ∆ρgR4 is analogous to the Bond number. Experiments performed on pure POPC membranes have determined 0/κb [44] 7 that κb ranges from 10kBT to 40kBT [44, 45] where kB is the Boltzmann constant and T is absolute temperature. We have chosen the value of 40kBT as the characteristic value to scale the experimental data and as a parameter in the computer simulations. The traction jump ∆f ≡ (σ(1) − σ(2)) · n, where n is the unit vector normal to the surface, is determined from the configuration of the membrane. This term will be further discussed in the following section. The Green's function kernels G and T for a wall bounded semi-infinite space were derived by Blake [46] and can be viewed as the fundamental solution to Stokes equations plus additional terms to account for the presence of the planar wall. Membrane mechanics The traction jump ∆f introduced in Eq. 3 must include all external forces acting on the vesicle. For the problem under consideration ∆f = ∆f mem + ∆f grav + ∆f elec, where the terms on the right refer to the contribution of the membrane elasticity, the gravitational pull and the electrostatic interactions, respectively. Vesicle deformation at equilibrium results from the balance of all these terms. There has been extensive theoretical research on lipid membrane deformation that involves the min- imization of the free energy functional first proposed by Helfrich [7]. In general, it has been observed that the shape of a vesicle in suspension is limited by its resistance to bending, which is governed by the bending rigidity modulus of the membrane. The shape of the vesicle is determined by minimization of the shape energy which may be written as HH = κb 2 (cid:73) (cid:73) dS, (C1 + C2)2dS + σ (4) where C1 and C2 are the principal curvatures of the membrane. Both integrals are over the surface area of the membrane. The last term in Eq. 4 takes into account the constraint of constant area [35]. It is customary to assume membrane incompressibility based on the high energy cost of surface area dilation of the lipid membrane as compared to other modes of deformation. Most authors rely on approximations based on the calculation of membrane tensions that are proportional to membrane strain [10, 14] which limit deformability of the membrane. Incompressibility can be imposed exactly by explicitly calculating the tension corresponding to the Lagrange multiplier by enforcing a divergence free velocity field on the membrane surface. This has been done for vesicles undergoing small deformations [23, 8 47,48] and for vesicles in a wall-bound shear flow [16]. An algorithm for the simulation of three dimensional vesicle dynamics was developed by Boedec and coworkers [15]. Micropipette aspiration experiments have shown that vesicle strain can occur at low tensions [34]. It has been hypothesized that this strain in the low-tension regime results from the smoothing of thermal membrane fluctuations occurring in the suboptical range, which act as an area reservoir. Area dilation as a function of tension is thus given by [34] α = kBT 8πκb ln (1 + cΣA) + Σ Kext . (5) where c is a coefficient that depends on the topology of the vesicle, Σ = σR2 0/κb is the effective reduced tension [44] and Kext = KAR2 0/κb is the dimensionless membrane's modulus for direct area expansion. The variable A is the dimensionless surface area of the membrane. Given the relatively large values of the membrane dilation modulus KA compared to κb, Kext is a large number of order 108. At low tensions, stretching of the membrane is ruled by the thermal energy (logarithmic term in Eq. 5) which is characterized by the bending coefficient, while at high tensions membrane strain involves surface dilation and is dominated by Kext (linear term in Eq. 5). While membrane strain due to direct surface dilation is difficult to achieve under normal flow conditions due to the relatively high magnitude of Kext, smoothing of thermal undulations are easily accessible at the low tensions that arise during gravity induced sedimentation. In the present work we use Eq. 5 to calculate the tension on the membrane for a given area strain. Although we are modeling a dynamic process, the sedimentation is slow and will be regarded as a quasi-equilibrium process for which a uniform membrane tension can be assumed. We have performed simulations which have included the calculation of local tensions to enforce local incompressibility and have found the tangential stresses due to tension gradients to be at least two orders of magnitude smaller than the other forces acting on the membrane at any given time for the conditions being considered. The contribution of the membrane elasticity to the force density can be obtained through the functional derivative of Eq. 4 as given by [35]. In non-dimensional form the Helfrich force can be expressed as ∆f mem = −(cid:0)(cid:0)4H 3 + 2∇2 sH − 4KGH(cid:1) + 2ΣH(cid:1) n. (6) Here, H = (C1 + C2)/2 is the mean curvature, KG = C1C2 is the Gaussian curvature and ∇2 s is the 9 Laplace-Beltrami operator over the surface of the membrane. Since the contribution of the curvature term to the total energy will depend on the bending modulus, membrane deformation will depend on factors that influence this parameter such as lipid composition [45]. The gravitational pull is incorporated into the calculations through [30] ∆f grav = −g0z, (7) where z is the vertical position of each differential element of the membrane with respect to a reference plane, in this case the glass surface, which has been made dimensionless by scaling with respect to R0. The calculation of the electrostatic contribution ∆f elec is explained in the following section. Electrostatic interactions Experiments performed in the absence of salt result in long screening lengths. For this reason the electrostatic interaction of the vesicle with the surface will play an important role in determining the sedimentation dynamics of the vesicle and the equilibrium state. Microscope glass slides were used as substrates in the experiments. The glass, unless treated using methods such as silanization [49], will present a surface charge, where the main mechanism by which the glass surface acquires charge while in contact with water is by dissociation of the silanol groups [50]. The glass surface charge density is negative and has a value around σglass = −0.2 mC/m2 for glass in water at pH 7.5 [51, 52]. Vesicles are labeled with 1 mol% NBD-PE fluorescent probe, which has a negative charge and gives the membrane a negative surface charge density [53]. Hence, the electrostatic interaction between the vesicle and the glass will be repulsive. We can obtain an expression for the interaction force between the membrane and the glass surface from the linearized Poisson-Boltzmann (P-B) equation [36, 37] Ψ is the electric potential and κ−1 is the Debye length: (cid:19) 1 2 , κ = ∇2Ψ = κ2Ψ. (cid:18) 2e2C εwkBT (8) (9) 10 where e is the fundamental charge, C is the ion concentration and εw is the dielectric constant of water. Equation 8 is valid for potentials less than 40 mV (about twice the thermal potential, kbT /e) [37]. In our first set of experiments no salts are added to the solutions, resulting in a Debye length of about 300 nm due to the presence of hydroniums and residual salts. When calcein is added to the solution the Debye length is lowered to 214 nm.The interaction between the vesicle and the glass surface can be modeled using the linear Derjaguin approximation which expresses that the interaction force between a small plane segment of the membrane and the plane surface of the glass is given by [36] f (h) = εw 8π κ2ΨvesΨglass cosh (κh/2) (10) where h is the distance between the two surfaces and the potentials, for the glass and the vesicle are given, respectively, by: and Ψglass = σglass εwκ Ψves = σvesR0 (1 + κR0)εw . (11) (12) The surface charge density of the vesicle can be estimated to be around σves ∼ −2.9 mC/m2 and its size ranges from 1 µm to 50 µm in radius. Thus, the potential generated by the vesicle is high enough that the linearized P-B Eq. 8 cannot accurately model its behavior. However, for distances greater than the Debye length the behavior of the solutions of the linear and non-linear forms are very similar, provided the surface potential, Ψves, is replaced by a renormalized surface potential given by [54, 55] Ψ∗ ves = 4kBT e . (13) which is independent of σves and has a value of around 100 mV in our experimental setup. Substituting Eq. 13 into 10 for the vesicle potential, the expression to calculate the electrostatic repulsion becomes f (h) = kBT 2πe Ψglassκ2εw cosh (κh/2) (14) Combining Eqs. 11 and 14 and expressing in terms of non-dimensional parameters, the contribution of the electrostatic interaction yields: ∆f elec = ψie cosh (κie z h/2) Where we have introduced the dimensionless parameters κie = R0κ and ψie = R2 0κieσglasskBT 2πeκb 11 (15) (16) Note that in Eq. 15 the distance h = h/R0 is expressed in dimensionless form dividing the distance by the characteristic length R0. A total of six dimensionless parameters have been introduced, two from the fluid dynamics model, κ and g0, two from the electrostatic model, κie and ψie, and two from the membrane mechanics model, Kext and the reduced effective tension Σ. The latter depends on the instantaneous configuration of the membrane and is recalculated at each time step. When both the external and internal fluids have the same viscosity, κ = 0 which simplifies Eq. 3 by canceling the second integral in the right hand side. Given that we found a very small difference between the viscosities of the sucrose and glucose solutions, we used this simplifying assumption in the solution of the model. The remaining four dimensionless parameters are calculated in order to match the experimental con- ditions. The radius R0 of the vesicle together with all other physical parameters determine the values of κie, ψie, Kext and g0. A range of vesicle radii was chosen to match experimental condition. Numerical method The general boundary integral method has been widely used to model the dynamic behavior of drops, capsules, vesicles and cells suspended in general flows. We refer to [31,43,56] for the general description of the method. However, in the following paragraphs we discuss particular considerations for the simulation of gravity-induced sedimentation. To obtain the computational domain we use the method of uniform triangulation to discretize the initially spherical membrane into a uniform triangulated mesh. As it has been described by other au- thors [31] we begin with a regular icosahedron inscribed into the sphere. Each face of the icosahedron is 12 divided into four smaller triangles by dividing each edge at its midpoint, the new vertices are projected radially onto the sphere and the process is applied recursively as many times as necessary to obtain the desired refinement. In the current work we do three refinements to obtain 642 vertices and 1280 triangular elements. An example of the computational mesh is shown in Fig. 2. Integration of the right hand side terms in Eq. 3 are approximated by a simple surface trapezoidal rule that requires the integrands only at the triangle vertices [31]. This rule can be written, for any function g(x), in the following form, by reassigning the contribution of each triangular element to the vertex: (cid:90) g(x) ≈(cid:88) (cid:88) i ∆Si = 1 3 S, g(x)∆Si, (17) (18) (19) (20) where the summation in (18) is over all flat triangular areas ∆S with vertex xi. One of the greatest challenges in the numerical algorithm is the calculation of the curvatures and the Laplace-Beltrami operator in Eq. 6 over the discrete triangulated surface. A common algorithm used to calculate the mean and Gaussian curvature is to fit via least-squares a quadratic surface to each node and its neighbors and calculate the curvature from this function. From our tests we have determined that this method introduces sufficient error in the calculation of the curvature to be problematic during the calculation of the Laplace-Beltrami operator. For this reason we have decided to use the discrete differential-geometry operators as presented in [57]. The mean curvature estimate is derived from a discretization of the Laplace-Beltrami operator applied to the 1-ring neighborhood. Given a patch of triangles surrounding point xi as shown in Fig. 3, the estimates for the Gaussian curvature, Ki and mean curvature Hi, at xi, given by Meyer et. al. [57] are: 2π −(cid:88)  θj j (cot αij + cot βij) (xi − xj) Ki = 1 AVoronoi 1 (cid:88) 2Hini = AVoronoi i where xj, θj, αij and βij are shown in the figure. The area AVoronoi corresponds to the "Voronoi area", defined in each triangle by the point xi, the midpoints of the triangle edges, and the circumcenter of the triangle, summed over all the triangles. Meyer et. al. demonstrated that the error in the computation of the curvature is minimized by using this area, as opposed to other alternatives. The Voronoi area can be calculated by: AVoronoi = 1 8 (cot αij + cot βij)(cid:107)xi − xj(cid:107)2 (cid:88) j (cid:88) 13 (21) (22) Finally the Laplace-Beltrami operator of the mean curvature H can be estimated as [57] ∇2 sH = 1 AVoronoi j∈N1(i) (cot αij + cot βij) (Hj − Hi) A similar method was recently used by Boedec and coworkers [15] for the calculation of the curvatures and Laplace-Beltrami operator. The numerical algorithm advances in a relatively simple fashion and can be described as follows: from the shape of the vesicle the traction ∆f on the membrane is computed, the boundary-integral Eq. 3 can then be solved for the velocity at the nodes, finally nodes are advected via integration in time using a second order Runge-Kutta method. The algorithm continues until an equilibrium state is achieved. An important difficulty in the boundary-integral calculations is efficient mesh control. Namely, if the nodes are simply advected with the flow, an initially regular mesh of triangles covering the surface becomes highly irregular after a short simulation time, thus invalidating the calculation. Mesh degeneration is especially problematic in gravity-induced motion given the large displacements of the vesicles. A mesh stabilization method can be developed by, instead of advecting the membrane nodes with the interfacial velocity u from Eq. 3, using the normal velocity (u · n)n plus an artificial tangential velocity field which can be constructed to maintain certain uniformity of the mesh. In the current work we construct the tangential velocity field using the passive mesh stabilization algorithm first introduced by Zinchenko and coworkers [31]. The algorithm is based on the global minimization of the rate of change of the distances between neighboring nodes. To achieve this one must solve an optimization problem over the whole surface of the membrane. We found that by using such algorithm long simulations of gravity-induced sedimentation of vesicles were possible with reasonable time steps. Results and Discussion In this section the results from the computer simulations are presented and compared to the experimental data. First, the sedimentation dynamics are studied, then the equilibrium state of the vesicle is analyzed. 14 Sedimentation of vesicles towards a flat surface The sedimentation of vesicles in isotonic conditions was studied using SPIM. The sedimentation process is driven by the density difference between the sucrose and glucose solutions (≈35 kg/m3). The best six image series were analyzed. The criterion for this selection was to ensure that the vesicle did not present significant lateral displacement during sedimentation. Fig. 4 shows sample images from one of the series that were analyzed. The sedimentation rate from the images is calculated by a simple two point backwards finite difference scheme for the position of the centroid between successive images. All experimental results have a time lapse between images of 15 seconds. Due to the optical configuration of the SPIM, an interference pattern results in an image artifact where some regions of the membrane appear to be brighter. This could be interpreted as these regions having a higher concentration of fluorescente probe. Nevertheless, given the preparation technique and the experimental configuration we do not expect this to be the case. We consider the assumption of uniformly distributed probe to be more accurate. The figure also shows the sedimentation sequence as predicted by our computer simulations under similar conditions. Computationally, we calculate the sedimentation rate with a method similar to the one used experimentally, but the time between data points is much smaller. The simulations are initialized with the vesicle at a certain distance from the surface (between 6-8 radii). The position of the centroid of the vesicle is calculated at each time step and the sedimentation rate is calculated. Sedimentation rates calculated experimentally and computationally are shown in Fig. 5. In the figure the distance to the surface is set to non-dimensional units by dividing by R0 and velocity is also given in non-dimensional form. Far away from the surface, the simulations match the experimental results accurately. The sedimentation velocity decreases as the vesicle approaches the flat surface and the rate of change is well predicted by the simulations. At distances larger than 0.2 the electrostatic interactions are negligible and the whole process is dominated by the hydrodynamics. As the vesicle approaches the wall, it decelerates. For uncharged vesicles (dark solid line in the figure) the velocity decreases exponentially approaching zero. This behavior is a result of the slow draining of the fluid that separates the vesicle and the surface, which corresponds to a lubrication regime [29]. When the electrostatic interaction is taken into account (light solid line in the figure) the sedimentation velocity decreases rapidly to zero further away from the surface avoiding this regime entirely. Close to the wall the electrostatic interactions become increasingly important causing a rapid deceleration, as the 15 gravity-induced force is balanced by the electrostatic interactions. Closer to the surface there appears to be a larger discrepancy between the experimental and computa- tional results. Experimental data lies between the simulations with and without electrostatic interactions. We attribute this to the method used in determining the distance between the vesicle and the surface in the SPIM experiments, in which the gap between the vesicle and the surface is determined by calculating the distance between the centroids of the vesicle and its reflection (see Fig. 4) divided by two, minus the radius of the vesicle. This method does not take into account the deformation of the lower part of the vesicle, which induces an underestimation of the size of the gap at equilibrium. One could also look for an explanation to this discrepancy in the assumptions made when developing our model. The most important simplification affecting the sedimentation rate is the assumption of having a uniform tension on the membrane. Given the symmetry of the problem, during the sedimentation process the membrane will become immobile and flow inside of the vesicle will be restricted. With a uniform tension, the membrane is mobile and there is the possibility of flow inside of the vesicle, as it happens in sedimenting droplets. This phenomenon will directly affect sedimentation rate. We have performed simulations, and compared the velocity of particles with mobile and immobilized interfaces under the same conditions and have found differences of less than 0.1 percent under these conditions. This difference is well below our experimental uncertainty. Also, this effect would be more important far away from the wall where the sedimentation rate is higher, opposite to our observations. Finally, in the following section we show that our model accurately predicts the equilibrium position of the vesicle as measured by the more precise confocal microscopy. Once the forces acting on the vesicle balance out, the vesicle reaches equilibrium and the sedimentation rate approaches zero. At this point the vesicle acquires a shape determined by the equilibrium of the electrostatic and gravitational forces with the membrane resistance to deform. Gap between the vesicle and the glass surface at equilibrium The gap between the vesicle and the surface at equilibrium was further studied using confocal microscopy images. Confocal images and the corresponding frame from the computer simulation for similar conditions are shown in Fig. 6. The small gap between the vesicle and the surface is seen for both the confocal image and the computer simulation. The gap at equilibrium measured experimentally and in the computer simulations are shown in Fig. 7 16 as a function of g0. The Debye length for NBD-PE labeled vesicles was calculated to be 214 nm when the ionic concentration of the calcein is accounted for. Computed simulations using that value for the Debye length show good agreement with the experimental results. The confocal microscopy combined with our image analysis algorithm provides a more accurate way of measuring the gap than the SPIM images. The distance to the surface decreases as g0 increases due to the increase in the effective weight of the membrane as compared to the electrostatic force. Experimentally, the addition of a small amount of salt results in the screening of the electrostatic forces and a decrease of the Debye length. By adding 3 mM of NaCl the Debye length is decreased to 5.5 nm. This condition can be used as a control, to verify that our observations are due to the electrostatic interactions. The experimental data and computer simulations under these conditions are also shown in Fig. 7. Overall, under these conditions, we observe a decrease of the equilibrium distance between the vesicle and the surface. The computer simulations predict a smaller gap distance than those measured experimentally, especially for smaller vesicles. Before commenting on the possible explanations for this behavior we note in passing that for larger vesicles the gap distance seems to converge for both Debye lengths being considered. This indicates that for heavier vesicles, which can reach closer to the wall, sedimentation is limited by the viscous draining of the fluid between the vesicle and the wall rather than the electrostatic interactions. The gap size is determined through image analysis using a custom ImageJ [58] algorithm. The process consists in plotting the intensity profile along a line that crosses the water gap between the vesicle and the glass surface. From the plot, the gap size is defined as the distance between points with an intensity corresponding to 70 percent of the highest measured value. Several distances measured for the same vesicle (between 3 and 5) are then averaged. The error bars shown in figures 7 and 8 are obtained from a combination of the deviation from the mean of the measured values with a confidence interval of 95%, together with the bias error due to the resolution of the microscope configuration used (around 0.29 microns). Our scaling normalizes the weight of all vesicles. It can be observed in Fig. 5 that sedimentation rate is independent of the size of the vesicle in non-dimensional terms. Physically, sedimentation rate does depend on the size of the vesicle, as heavier vesicles sediment faster. From the characteristic length R0 and the characteristic time defined in Eq. 2 it can be seen that physical velocity varies with R2 0. That is, vesicles with radius of order one are sedimenting a hundred times slower that those of radius 10! Due to 17 the stability of the vesicles, experiments are limited to one hour. With this in mind, a better explanation to the observed behavior is that smaller vesicles have not yet reached equilibrium and are sedimenting very slowly when the measurement is performed. Computationally, we are allowing the vesicles to reach equilibrium defined as the state when displacements are below the numerical resolution of our algorithm, thus the measured gap is smaller. Overall, the computational model is able to predict in an accurate fashion the size of the gap at equilibrium for a wide variety of conditions. The agreement in the results also indicates an appropriate selection of the values of some important parameters that were not measured in our experiments but relied on calculated values and theoretical predictions, such as the glass surface charge density and the counterions concentration in the solutions. So far we have presented results corresponding to charged vesicles labeled with NBD-PE in solution. We have considered solutions with no added salt where electrostatic interactions are important (relatively large Debye length) and with salt added to screen the effect of electrostatic interactions. The addition of a small amount of salt (3 mM) generates a slightly hypertonic condition which causes the vesicle to deflate. In order to investigate the effect of this on the measured gap we perform additional experiments with unlabeled vesicles (no charge). The gap distance for charged vesicles in a solution with added salt compared to unlabeled vesicles is shown in Fig. 8. It can be seen that there is no difference in the behavior of both groups and that the same behavior discussed above is observed in unlabeled vesicles. In the following section we investigate the deformation of both charged and uncharged vesicles. The thorough understanding of the role played by the electrostatic interactions on the sedimentation process, as well as on determining its position at equilibrium might be important in understanding the formation and stability of certain colloidal systems. The results presented in this section will play an important role in this understanding. Vesicle deformation at equilibrium When the vesicle reaches equilibrium, the sedimentation process stops and the vesicle adopts an equilib- rium shape determined by the balance of the forces acting on it. The shape of the vesicle at equilibrium was studied through confocal microscopy and image analysis as explained in the Materials and Methods section. The analysis was applied to images similar to that shown in Fig. 6 (A). Contrast was improved by staining the glucose solution with low concentrations of calcein (1 µM). At these low concentrations, 18 it should not influence the osmotic conditions of the solution. It has been shown [34] that membrane deformation is caused by the superposition of two modes: smoothing of thermal undulations and direct stretching of the membrane. At low external forces, such as those in the present experiments, it is expected that strain be due to the former mode only. Tension on the membrane at equilibrium is proportional to the value of g0. By plotting the area strain as a function of g0 in a semi-log plot (Fig. 9) it is evident that the logarithmic term in Eq. 5 is the dominating term in defining the shape of the membrane at equilibrium as expected. In Fig. 6 experimental images of deformed vesicles at equilibrium were shown together with the compu- tational prediction at similar conditions. It is interesting to note that those relatively large deformations are achieve with very small area strains (less than 0.01 in all cases), which can be explained due to the smoothing of thermal undulations [33]. Hence, it becomes apparent that considerations of thermal undulations is important in the study of vesicle sedimentation. The tension–strain relationship (Eq. 5) suggests that strain at a certain tension depends on the mem- brane bending rigidity. It has been suggested previously [59–61] that electrostatic interactions between neighboring NBD-PE molecules should provide a contribution to the bending rigidity. In order to verify this observation we compare the strain of charged and uncharged vesicles. As mentioned above, salt produces a slightly hypertonic condition which causes the vesicle to deflate, this will certainly affect the deformability of the particles. For this reason in the following we report the deformation of both labeled and unlabeled vesicles suspended in the same solution with no salt added. In the previous section it was shown that not labeling the vesicles had the same effect on the measured gap as adding a small amount of salt to screen the electrostatic interactions. Fig. 9 shows the measured area strain for both NBD-PE labeled and unlabeled vesicles. The simulation results for vesicles with a bending rigidity modulus κb = 40kBT is also shown in the figure. For the majority of the unlabeled vesicles the measured strain is in line with the prediction of the computational model. For a smaller fraction of the data points, mostly lighter vesicles, a much smaller strain is measured. All data points correspond to the same sample, hence tonicity effects should be ruled out. The most probable cause for this observation is, as explained above, that this vesicles have not yet reached an equilibrium state and have not yet fully deform to balance the gravitational force. Charged vesicles undergo a smaller strain at equilibrium. Solid triangular data points in Figs. 7 and 9 correspond to the same vesicles. Based on the comparison between the experimental data and the 19 computer simulations, we consider that these vesicles have reached equilibrium, and electrostatic forces are balanced with the gravitational force. These results suggest that the presence of charged fluorescent probes does have the effect of increasing the bending rigidity as has been reported previously [59]. Measurement of the mechanical properties of vesicles represents one of the possibilities presented by the current computational model. By comparing good experimental data with the computer simulations, one might be able to determine the mechanical properties of membranes which could complement techniques such as micropipette aspiration. Conclusions We have reported the implementation of a computational algorithm for the simulation of lipid vesicles in suspension. Experiments using SPIM and confocal microscopy were performed to verify the performance of the algorithm. The algorithm is used to simulate the sedimentation of vesicles due to gravity, taking into consideration the electrostatic interactions between the membrane and a glass surface towards which it is sedimenting. We have shown that our simulations are able to predict, with reasonable accuracy, the rate at which the vesicles sediment and the velocity variations as the vesicle approach the glass surface. The algorithm also predicts the equilibrium gap between the vesicle and the surface at equilibrium. It was shown that the consideration of the electrostatic contribution to surface interactions is essential in order to accurately predict the sedimentation rate, especially at close range from the surface, and the fluid gap between the vesicle and the surface at equilibrium. We have shown that by modeling the mechanical behavior of the membrane with Eq. 5, which su- perposes the deformation due to the smoothing of thermal undulations and direct membrane stretch, area strains are reasonably small and in line with those observed experimentally. The model used in our simulations also takes into account the intrinsic mechanical properties of the membrane through the membrane bending modulus, which can be modulated in cells through membrane composition. For example, by adding 30% cholesterol the bending rigidity of the membrane can be doubled [62]. It is also suggested that charged fluorescent probes on the membrane have the effect of increasing its bending rigidity. Our vesicle model has considered the basic physical contributions: Stokes flow, bending rigidity, membrane deformation through smoothing of thermal undulations but limited by the high energy cost 20 of direct membrane deformation and electrostatic interactions between the vesicle and the glass surface. In its current form the algorithm can be used to study the effect that the different physical parameters discussed have on the sedimentation rate, distance to the surface at equilibrium and membrane strain. It can also be extended in order to study important phenomena such as colloidal stability, and biofilm formation. Finally, regarding future work, we consider our current model to be robust and accurate enough to allow us to investigate the roll that the physics described and studied in the current paper play in the slow motion of a suspended vesicle. Nevertheless, the dynamics and motion of vesicles under shear or general flow pose increasing challenges as the calculation of local tensions becomes mandatory. The method proposed by Boedec et al. is a very good starting point, but the inclusion of viscosity contrasts different than one and other physics, such as those investigated in the current paper, could prove challenging and to the best of our knowledge remain unsolved. Acknowledgments We acknowledge Brice Ronsin from the Toulouse RIO Imagin facility (CNRS UMR 5547 Universit´e Paul Sabatier, Toulouse) and the MEMPHYS group from Southern Denmark University for their help in the development of the experimental procedures of this work. References 1. Tomaiuolo G, Guido S (2011) Start-up shape dynamics of red blood cells in microcapillary flow. Microvascular research 82: 35–41. 2. De Rosa M, Alinovi C, Galtieri A, Scatena R, Giardina B (2007) The plasma membrane of ery- throcytes plays a fundamental role in the transport of oxygen, carbon dioxide and nitric oxide and in the maintenance of the reduced state of the heme iron. Gene 398: 162–171. 3. Yoon Y, Hong H, Brown A, Kim D, Kang D, et al. (2009) Flickering analysis of erythrocyte me- chanical properties: dependence on oxygenation level, cell shape, and hydration level. Biophysical journal 97: 1606–1615. 21 4. Vlahovska PM, Podgorski T, Misbah C (2009) Vesicles and red blood cells in flow: From individual dynamics to rheology. Comptes Rendus Physique 10: 775–789. 5. Kaoui B, Farutin A, Misbah C (2009) Vesicles under simple shear flow: Elucidating the role of relevant control parameters. Physical Review E 80: 61905. 6. Kaoui B, Biros G, Misbah C (2009) Why do red blood cells have asymmetric shapes even in a symmetric flow? Physical review letters 103: 188101. 7. Helfrich W (1973) Elastic properties of lipid bilayers: theory and possible experiments. Z Natur- forsch 28: 693–703. 8. Seifert U (1995) The concept of effective tension for fluctuating vesicles. Zeitschrift Fur Physik B 97: 299–309. 9. Seifert U (1997) Configurations of fluid membranes and vesicles. Advances in Physics 46: 13–137. 10. Biben T, Farutin A, Misbah C (2011) Three-dimensional vesicles under shear flow: Numerical study of dynamics and phase diagram. Physical Review E 83: 31921. 11. Sukumaran S, Seifert U (2001) Influence of shear flow on vesicles near a wall: A numerical study. Physical Review E 6401. 12. Misbah C (2006) Vacillating breathing and tumbling of vesicles under shear flow. Physical Review Letters 96: 28104. 13. Kaoui B, Ristow GH, Cantat I, Misbah C, Zimmermann W (2008) Lateral migration of a two- dimensional vesicle in unbounded poiseuille flow. Physical Review E 77: 21903. 14. Kaoui B, Harting J, Misbah C (2011) Two-dimensional vesicle dynamics under shear flow: Effect of confinement. Physical Review E 83: 066319. 15. Boedec G, Leonetti M, Jaeger M (2011) 3d vesicle dynamics simulations with a linearly triangulated surface. Journal of Computational Physics 230: 1020–1034. 16. Zhao H, Spann A, Shaqfeh E (2011) The dynamics of a vesicle in a wall-bound shear flow. Physics of Fluids 23: 121901–121901. 22 17. Abkarian M, Viallat A (2005) Dynamics of vesicles in a wall-bounded shear flow. Biophysical Journal 89: 1055–1066. 18. Kantsler V, Steinberg V (2005) Orientation and dynamics of a vesicle in tank-treading motion in shear flow. Physical Review Letters 95: 258101. 19. Kantsler V, Steinberg V (2006) Transition to tumbling and two regimes of tumbling motion of a vesicle in shear flow. Physical Review Letters 96: 36001. 20. Deschamps J, Kantsler V, Steinberg V (2009) Phase diagram of single vesicle dynamical states in shear flow. Physical review letters 102: 118105. 21. Veerapaneni SK, Rahimian A, Biros G, Zorin D (2011) A fast algorithm for simulating vesicle flows in three dimensions. Journal of Computational Physics 230: 5610 - 5634. 22. Danker G, Vlahovska P (2009) Vesicles in poiseuille flow. Physical review letters 102: 148102. 23. Boedec G, Jaeger M, Leonetti M (2012) Settling of a vesicle in the limit of quasispherical shapes. Journal of Fluid Mechanics 690: 227–261. 24. Huang Z, Abkarian M, Viallat A (2011) Sedimentation of vesicles: from pear-like shapes to mi- crotether extrusion. New Journal of Physics 13: 035026. 25. Kushner J, Rother M, Davis R (2001) Buoyancy-driven interactions of viscous drops with deforming interfaces. Journal of Fluid Mechanics 446: 253–269. 26. Rother M, Davis R (2004) Buoyancy-driven coalescence of spherical drops covered with incom- pressible surfactant at arbitrary p´eclet number. Journal of Colloids and Interface Science 270: 205–220. 27. Rother MA, Zinchenko AZ, Davis RH (2006) Surfactant effects on buoyancy-driven viscous inter- actions of deformable drops. Colloids and Surfaces A: Physicochemical and Engineering Aspects 282: 50–60. 28. Rother MA, Davis RH (2008) Buoyancy-driven breakup of an isolated drop with surfactant. Phys- ical Review Letters 101: 44501. 23 29. Ascoli EP, Dandy DS, Leal LG (1990) Buoyancy-driven motion of a deformable drop toward a planar wall at low reynolds number. Journal of Fluid Mechanics 213: 287–311. 30. Pozrikidis C (1992) The buoyancy-driven motion of a train of viscous drops within a cylindrical tube. Journal of Fluid Mechanics 237: 627–648. 31. Zinchenko A, Rother M, Davis R (1997) A novel boundary-integral algorithm for viscous interaction of deformable drops. Physics of Fluids 9: 1493. 32. Zinchenko AZ, Rother MA, Davis RH (1999) Cusping, capture, and breakup of interacting drops by a curvatureless boundary-integral algorithm. Journal of Fluid Mechanics 391: 249–292. 33. Ly HV, Longo ML (2004) The influence of short-chain alcohols on interfacial tension, mechanical properties, area/molecule, and permeability of fluid lipid bilayers. Biophysical Journal 87: 1013– 1033. 34. Evans E, Rawicz W (1990) Entropy-driven tension and bending elasticity in condensed-fluid mem- branes. Physical Review Letters 64: 2094–2097. 35. Zhong-Can OY, Helfrich W (1989) Bending energy of vesicle membranes: General expressions for the first, second, and third variation of the shape energy and applications to spheres and cylinders. Physical Review A 39: 5280–5288. 36. Derjaguin B (1940) On the repulsive forces between charged colloid particles and on the theory of slow coagulation and stability of lyophobe sols. Transactions of the Faraday Society 35: 203–215. 37. Carnie S, Chan D, Gunning J (1994) Electrical double layer interaction between dissimilar spherical colloidal particles and between a sphere and a plate: The linearized poisson-boltzmann theory. Langmuir 10: 2993–3009. 38. Stankovich J, Carnie S (1996) Electrical double layer interaction between dissimilar spherical col- loidal particles and between a sphere and a plate: Nonlinear poisson-boltzmann theory. Langmuir 12: 1453–1461. 39. Menger F, Angelova M (1998) Giant vesicles: imitating the cytological processes of cell membranes. Accounts of Chemical Research 31: 789–797. 24 40. Greger K, Swoger J, Stelzer E (2007) Basic building units and properties of a fluorescence single plane illumination microscope. Review of Scientific Instruments 78: 023705. 41. Bradski G (2000) The opencv library. Doctor Dobbs Journal 25: 120–126. 42. Nelder J, Mead R (1965) A simplex method for function minimization. The Computer Journal 7: 308. 43. Pozrikidis C (1992) Boundary integral and singularity methods for linearized viscous flow. New York: Cambridge University Press, first edition. 44. Henriksen J, Rowat A, Ipsen J (2004) Vesicle fluctuation analysis of the effects of sterols on mem- brane bending rigidity. European Biophysics Journal 33: 732–741. 45. Olsen B, Schlesinger P, Baker N (2009) Perturbations of membrane structure by cholesterol and cholesterol derivatives are determined by sterol orientation. Journal of the American Chemical Society 131: 4854–4865. 46. Blake J (1971) A note on the image system for a stokeslet in a no-slip boundary. In: Proc. Camb. Phil. Soc. Cambridge Univ Press, volume 70, pp. 303–310. 47. Seifert U (1999) Fluid membranes in hydrodynamic flow fields: Formalism and an application to fluctuating quasispherical vesicles in shear flow. The European Physical Journal B-Condensed Matter and Complex Systems 8: 405–415. 48. Zhao H, Shaqfeh E (2011) The dynamics of a vesicle in simple shear flow. Journal of Fluid Mechanics 674: 578–604. 49. Cohen J (2003) Electrophoretic characterization of liposomes. Methods in enzymology 367: 148– 176. 50. Iler R (1979) The chemistry of silica: solubility, polymerization, colloid and surface properties, and biochemistry. Wiley New York. 51. Polin M, Grier D, Han Y (2007) Colloidal electrostatic interactions near a conducting surface. Physical Review E 76: 041406. 25 52. Behrens S, Grier D (2001) The charge of glass and silica surfaces. The Journal of Chemical Physics 115: 6716. 53. Groves J, Boxer S, McConnell H (1998) Electric field-induced critical demixing in lipid bilayer membranes. Proceedings of the National Academy of Sciences 95: 935. 54. Trizac E, Bocquet L, Aubouy M (2002) Simple approach for charge renormalization in highly charged macroions. Physical Review Letters 89: 248301. 55. Tellez G (2011) Nonlinear screening of charged macromolecules. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences 369: 322–334. 56. Pozrikidis C (2001) Interfacial dynamics for stokes flow. Journal of Computational Physics 169: 250–301. 57. Meyer M, Desbrun M, Schroder P, Barr AH (2002) Discrete differential-geometry operators for triangulated 2-manifolds. Visualization and mathematics 3: 34–57. 58. Abramoff M, Magalhaes P, Ram S (2004) Image processing with imagej. Biophotonics International 11: 36–42. 59. Rowat AC, Hansen PL, Ipsen JH (2004) Experimental evidence of the electrostatic contribution to membrane bending rigidity. EPL (Europhysics Letters) 67: 144. 60. Mitchell D, Ninham B (1989) Curvature elasticity of charged membranes. Langmuir 5: 1121–1123. 61. Winterhalter M, Helfrich W (1988) Effect of surface charge on the curvature elasticity of mem- branes. The Journal of Physical Chemistry 92: 6865–6867. 62. Pan J, Mills T, Tristram-Nagle S, Nagle J (2008) Cholesterol perturbs lipid bilayers nonuniversally. Physical Review Letters 100: 198103. Figure Legends 26 Figure 1. Schematic of vesicle subjected to gravity induced sedimentation. Sedimentation is driven by the density difference ∆ρ = ρ1 − ρ2. 27 Figure 2. Example of the mesh used in the initial configuration. The mesh is constructed by refining the triangular faces of an icosahedron inscribed into a sphere by dividing recursively each triangle into four smaller triangles and projecting the resulting nodes to the surface of the sphere. The mesh used in the current work is the result of four refinements in order to obtain a total of 642 nodes and 1280 triangular elements. Figure 3. 1-ring neighborhood of vertex indicating the sub-area used for computation using the method of Meyer et. al. 28 Figure 4. Image sequence for the sedimentation of a vesicle. The time interval between images is 15 s. A. Images obtained through SPIM for g0 = 3. B. Images from computer simulation depicting sedimentation of a vesicle under similar conditions. Figure 5. Sedimentation rate as a function of the distance to the surface. Experimental data (symbols) and computer simulations (lines) are shown. The light solid line corresponds to a simulation which incorporates the electrostatic interaction between the vesicle and the surface, the dark solid line corresponds to a simulation which does not take into account this interaction. 29 Figure 6. Images of vesicles at equilibrium A. and C. show confocal images of giant unillamellar vesicle at equilibrium for solutions with 3 mM of NaCl and no salt, respectively. Staining the outer solution with calcein (1.0 µM) allows to visualize the gap between the vesicle and the glass surface. B. and D. show computer simulation of a vesicle at equilibrium under similar conditions as in A. and C., respectively. Figure 7. Equilibrium distance to the surface. Gap at equilibrium as a function of g0 obtained from simulations (dashed lines) and experimental data (symbols) for two different Debye lengths (κ−1). In the simulations κ−1 is an input parameter, experimentally it depends on the ionic concentration which can be changed by the addition of salt. 30 Figure 8. Comparison between the effect of significantly screening the electrostatic interaction and absence of charge on the membrane. Open circles represent experimental data for the equilibrium distance between the vesicle and the surface for non-charged vesicles, while filled triangles show the distance to the surface at which charged vesicles reach equilibrium when 3 mM of salt is added. Figure 9. Membrane strain at equilibrium measured from experiments and simulations as a function of g0. Open circles and filled triangles show the measured strain for non-charged and charged vesicles, respectively. Computer simulation results are shows as filled squares and dashed line. The presence of charge in the membrane appears to stiffen the membrane resulting on smaller measured strains.
1905.04331
2
1905
2019-11-11T15:13:51
On the growth of non-motile bacteria colonies: an agent-based model for pattern formation
[ "physics.bio-ph" ]
In the growth of bacterial colonies, a great variety of complex patterns are observed in experiments, depending on external conditions and the bacterial species. Typically, existing models employ systems of reaction-diffusion equations or consist of growth processes based on rules, and are limited to a discrete lattice. In contrast, the two-dimensional model proposed here is an off-lattice simulation, where bacteria are modelled as rigid circles and nutrients are point-like, Brownian particles. Varying the nutrient diffusion and concentration, we simulate a wide range of morphologies compatible with experimental observations, from round and compact to extremely branched patterns. A scaling relationship is found between the number of cells in the interface and the total number of cells, with two characteristic regimes. These regimes correspond to the compact and branched patterns, which are exhibited for sufficiently small and large colonies, respectively. In addition, we characterise the screening effect observed in the structures by analysing the multifractal properties of the growth probability.
physics.bio-ph
physics
As part of the Springer Nature SharedIt initiative, a full-text view-only version of the paper can be accessed clicking here EPJ manuscript No. (will be inserted by the editor) 9 1 0 2 v o N 1 1 ] h p - o i b . s c i s y h p [ 2 v 1 3 3 4 0 . 5 0 9 1 : v i X r a On the growth of non-motile bacteria colonies: an agent-based model for pattern formation Lautaro Vassallo1, David Hansmann1, and Lidia A. Braunstein1,2 1 Departamento de Física, Facultad de Ciencias Exactas y Naturales, Universidad Nacional de Mar del Plata, and Instituto de Investigaciones Físicas de Mar del Plata (IFIMAR-CONICET), Deán Funes 3350, 7600 Mar del Plata, Argentina 2 Physics Department and Center for Polymer Studies, Boston University, Boston, Massachusetts 02215, USA the date of receipt and acceptance should be inserted later Abstract. In the growth of bacterial colonies, a great variety of complex patterns are observed in exper- iments, depending on external conditions and the bacterial species. Typically, existing models employ systems of reaction-diffusion equations or consist of growth processes based on rules, and are limited to a discrete lattice. In contrast, the two-dimensional model proposed here is an off-lattice simulation, where bacteria are modelled as rigid circles and nutrients are point-like, Brownian particles. Varying the nutri- ent diffusion and concentration, we simulate a wide range of morphologies compatible with experimental observations, from round and compact to extremely branched patterns. A scaling relationship is found between the number of cells in the interface and the total number of cells, with two characteristic regimes. These regimes correspond to the compact and branched patterns, which are exhibited for sufficiently small and large colonies, respectively. In addition, we characterise the screening effect observed in the structures by analysing the multifractal properties of the growth probability. PACS. XX.XX.XX No PACS code given 1 Introduction The concept of active matter is relatively new within soft matter physics; the fundamental units of this type of mat- ter, called active agents, have the particularity of absorb- ing energy from their environment and dissipating it in or- der to move, grow or replicate, among other activities [1]. Most of the examples of active matter are biological, such as bacteria. Although they can be seen as the simplest liv- ing organisms, they present interesting behaviours, both individually and collectively. Bacteria exhibit many different types of movement, de- pending on the species and the environment, which de- termines the macroscopic appearance of the colony. Ac- cording to Henrichsen [2], six types of motility can be identified: swimming, swarming, twitching, darting, glid- ing (newer studies subdivide this category [3]) and slid- ing. We focus on the last one, which is a mechanism pro- duced by the expansive forces of the colony, in combina- tion with special properties of the cell membranes charac- terised by low friction with the substrate on which they grow; the bacteria do not move by their own motors, but push each other by duplicating themselves and competing for the same spaces. Despite its simplicity, its importance has been pointed out in more complex bacterial processes, such as in the formation, dispersion and restructuring of biofilms [4]. Although we will not strictly enter the biofilms field in this work, it is important to mention they are the focus of numerous studies, due to the complexity of the processes involved, their resistance to hostile environ- ments, and the challenge they present to medicine [4 -- 8]. At the end of the 80s, Matsuyama [9], Fujikawa and Matsushita [10], showed that the patterns of bacterial colonies obtained in the laboratory could be fractal ob- jects. The properties of their patterns depend on two main factors: the concentration of nutrients, which in- fluences the growth rate of the colony, and the con- centration of agar, which determines the hardness of the substrate, and therefore, the mobility of the bac- teria. In the absence of special forms of motility, the patterns were classified in a two-dimensional phase di- agram in which five characteristic patterns were identi- fied: diffusion-limited aggregation-like (DLA-like), Eden- like, dense branching morphology (DBM), concentric ring and homogeneous disk-like. The experiments were per- formed mainly with the species Bacillus subtilis [10 -- 13]. Without self-propulsion, only DLA and Eden-like patterns are expected. At the theoretical level, continuous models are the most traditional and extended way of studying the pat- terns exhibited by bacteria colonies. In them, both bac- teria and nutrients (or any other variable of interest) are represented by density functions per unit area, and the spatio-temporal evolution of the system is described by systems of reaction-diffusion equations [11, 14 -- 23]. These models are successful in describing a wide range of pat- As part of the Springer Nature SharedIt initiative, a full-text view-only version of the paper can be accessed clicking here 2 Lautaro Vassallo et al: On the growth of non-motile bacteria colonies terns, although they are valid only at a mesoscopic scale. To represent growth at the microscopic level, microor- ganisms must be represented by discrete mobile entities (agents) [24 -- 27]. In this scheme, on-lattice [26] and off- lattice [24, 25, 27, 28] approaches can be chosen. The motivation of this work is to propose a micro- scopic model that can explain the experimental observa- tions, based on the fact that sliding is dominated by the mechanical interaction between the bacterial cells. As has been said, continuous models work well only on a meso- scopic scale, whereas in agent-based models, if space is discretised as Euclidean networks, mechanical laws can- not be used. In spite of being computationally expensive, in this work, we choose an off-lattice model, in order to represent our agents as rigid bodies governed by laws of mechanics. Thus, we can analyse the growth of bacterial colonies on a microscopic level. The off-lattice approach also avoids anisotropies in the patterns exhibited by the colonies induced by the discretization of the space. A typical way to characterise the complex structures that arise in surface growth is by means of the Hausdorff dimension, often referred to as the fractal dimension. How- ever, the fractal dimension is not a unique descriptor, as it was shown that two structures may have the same fractal dimension but are fundamentally different [29]. In order to describe structures more deeply and unequivocally, the de- termination of the multifractal properties of an associated measure (e.g. growth probability) offers a suitable supple- ment to the sole measurement of their fractal dimension. This is an entropy-based approach [30], classified this way to differentiate it from other analyses that rely only on metric concepts. Here the scaling properties are analysed for variations in different parts of the pattern, which are overlooked by a simple measurement of the fractal dimen- sion. As it is arduous to treat growth models with nonlocal rules analytically [31], it may be interesting to characterise an associated measure such as the growth probability to gain some insight about the process. A way to describe the multifractal behaviour is through the generalised dimensions Dq (also known as Rényi dimensions). If one covers the support of the mea- sure (set of all points where the measure is positive) with a set of boxes of size l and defines a probability Pi(l) (integrated measure) in the ith box, the generalised di- mensions Dq correspond to the scaling exponents for the i (l) ∼ l(q−1)Dq. In this P q context, q is typically referred to as the order q of the gen- eralised dimension Dq. Solving for Dq and taking the limit of l → 0, the conventional expression for the generalised dimensions is given by qth moments of Pi, defined by(cid:80) i ln(cid:80) Dq = 1 (q − 1) lim l→0 i ln l P q i (l) . For the case q = 1, the L'Hôpital's rule must be used; thus, (cid:80) Pi(l) log Pi(l) D1 = lim l→0 i ln l . The generalised dimensions are exponents that charac- terise the non-uniformity of the measure; the positive or- ders q accentuate the regions with higher probabilities while the negative q's the opposite. In early works, Dq was only defined for q ≥ 0 [32], so the previous definition had the particularity that the first value of the set, i.e. Dq=0, corresponded to the Hausdorff dimension of the support (because all the boxes have the same weight). There are also specific names for other cer- tain values. For example, Dq=1 is known as the informa- tion dimension, which is interesting in the case of diffusion- limited aggregations, since it can be physically interpreted as the fractal dimension of the active region, i.e., the un- screened region [29]; Dq=2 is known as the correlation dimension; Dq→±∞ are known as the Chebyshev dimen- sions, which are calculated with the maximum and min- imum probabilities, respectively; equivalences with other dimensions definitions can be made, even for fractional q values [30]. This is not the only multifractal analysis possible. In other works, the singularity spectrum is computed [33,34], which is closely related to the Rényi dimensions by a Leg- endre transform. Temporal fractals can also be studied, where the local scaling properties are now related to time behaviour [35]. We will use the generalised dimensions to characterise quantitatively the patterns produced by our model. 2 Model In this paper, we model the growth of non-motile bacterial colonies under different environmental conditions, specif- ically, nutrient concentration and nutrient diffusion. The growth rules are inspired by biology, as we capture the es- sential characteristics of bacteria without losing simplicity. We consider a two-dimensional and off-lattice space, which allows us to consider mechanical interactions between the agents, as we will explain below. There are two kinds of particles in the model, nutri- ent particles and bacterial cells. Both of them have phys- ical properties such as size, mass, position, velocity and might have applied forces. Nutrient particles are idealised as Brownian particles, so its initial velocities follow the Maxwell-Boltzmann distribution and evolve according to a Langevin equation of the form m v(t) = − κT D v(t) + f (t), where κ is the Boltzmann's constant, T is the tempera- ture and D is the diffusion coefficient. The function f (t) is a stochastic force whose components follow a Gaussian probability distribution with mean zero and standard de- viation σ = κT(cid:112)2/D. The Langevin equation is numer- ically integrated using a small time step ∆t, following the explanations in The Fokker-Planck Equation by H. Risken [36]. Nutrient particles are considered point-like, non-interacting with each other and with a small mass. The bacterial cells are modelled as rigid circles with radius rb, so they can interact with each other through normal forces. The numerical values used in the simulations can be found in Table 1. As part of the Springer Nature SharedIt initiative, a full-text view-only version of the paper can be accessed clicking here Lautaro Vassallo et al: On the growth of non-motile bacteria colonies 3 Table 1: Numerical values used for the simulation. Variable kT ∆t rb m (nutrient mass) Value [arbitrary units] 1 0.01 1 0.0001 In addition to the physical properties, bacterial cells have two biological characteristics: they can be fed and reproduce. The first one is an interaction with the nutri- ents, which are absorbed by the cell when they are in con- tact. Reproduction is the process by which the bacterium duplicates: an identical copy of the cell is generated in the same position as the original, so they overlap. Then, they are disaggregated by opposing velocities in a ran- dom direction. As a consequence, these cells may collide elastically with neighbours, according to the mechanics of rigid bodies, as shown in Fig. 1. When the cells stop over- lapping, they stop moving and become static because the medium is considered to be very viscous so that the mo- mentum gained by the collisions is immediately dissipated. All the calculations to resolve collisions and overlaps are based on an iterative constraint solver introduced by Erin Catto [37]. Fig. 1: Feeding and reproduction. Bacteria cells (circles) can absorb the particles diffusing in the medium (dots) and duplicate, causing collisions with neighbours, which move in the direction pointed by the arrows. The double blue arrow (colour online) points the direction in which the newborn bacteria disaggregate. In summary, reproduction causes both the movement of newborn cells and their neighbours, representing sliding motility. The simulation begins with a single cell in the origin of coordinates in a 2D substrate and a given quantity of nu- trient particles diffusing in space, according to the concen- tration and diffusion coefficient specified. When a nutrient particle touches the cell, it is absorbed and the bacteria duplicates. Now the colony is formed by two cells, which can absorb nutrients and reproduce. The process contin- ues in this way and the colony grows progressively. All the bacteria have a time delay (20 integration time steps), during which they can't duplicate; this rule ensures that Fig. 2: A ring surrounding the colony (in red) acts as a nu- trient reservoir. The concentration of nutrient is constant inside it, but decreases in the proximity of the colony, be- cause of the feeding. The distance between the ring and the colony is 60rb at least. no duplication occurs while newborn cells are still over- lapping and it is consistent with biological observations, e.g., Bacillus subtilis species has a delay of ∼ 25 minutes between duplications [17]. In order to keep the nutrient concentration constant, there is a ring that acts as a nutrient reservoir located at a given distance from the most external position of the bac- teria. This distance increases progressively as the colony grows, so the separation is 60rb at least. The nutrient con- centration within the ring (r > 60rb) is kept constant at a specified value. But, closer to the colony (r < 60rb), the concentration drops due to nutrient absorption by the bacteria. Periodic boundaries conditions are considered for the outer side of the ring (Fig. 2). The growth stops when the colony reaches a radius of 600rb, when characteristic patterns are fully developed. This implies that we have up to half a million cells forming the colony, depending on the parameters. 3 Results and discussions 3.1 First characterization of the structures In order to see the variety of morphologies that the model can produce, we choose several different values of nutrient concentration and nutrient diffusion, and register the posi- tion of the bacteria along the perimeter of the colony over time and average over a hundred realizations for each set of parameters. A morphology diagram is shown in Fig. 3a, where it can be seen that it is possible to generate round and compact colonies, as well as ramified, going through a variety of intermediate patterns. Similar morphological crossover can be seen in Figs. 3b -- e, which corresponds to the experiments carried out in [11,12]. The fractal dimen- sion of intermediate patterns in experiments was reported for the case of the most ramified one. In Table 2, we sum- marise the results found in the bibliography [10, 12] and ours (for the most ramified cases), which are in good agree- ment. As part of the Springer Nature SharedIt initiative, a full-text view-only version of the paper can be accessed clicking here 4 Lautaro Vassallo et al: On the growth of non-motile bacteria colonies (a) (b) (c) (d) (e) Fig. 3: Examples of different morphologies. In (a) there is a sample of the model predictions. On the X-axis, the diffusion coefficient D is varied (in arbitrary units), while on the Y -axis the nutrient concentration is varied (measured as the number of particles per unit area). Each curve, in a different shade of grey, corresponds to a different time. All other sub-figures correspond to experiments. The parameter Ca corresponds to the agar concentration, which determines the hardness of the substrate, and Cn corresponds to the nutrient concentration. (b) Ca = 10g/l; Cn = 20g/l. (c) Ca = 9g/l; Cn = 4.5g/l. (d) Ca = 8g/l; Cn = 3g/l. (e) Ca = 9g/l; Cn = 1g/l. Figure (b) is from [11], reprinted with permission from Elsevier. Figures (c), (d) and (e) are from [12], ©(1992) The Physical Society of Japan, reproduced with permission. Table 2: Fractal dimension Df . Values reported in exper- imental works, our model and DLA (the error is the stan- dard deviation). Only the results for the most branched cases are included with C = 1/80, D = 512 and C = 1/40, D = 512. Greater values than for the case of DLA are expected, where only one particle diffuses at a time, unlike our case where we have many (C > 0). Model [C = 1/80; D = 512] Model [C = 1/40; D = 512] Experiment [10] Experiment [12] DLA [29, 31] Df 1.760 ± 0.004 1.778 ± 0.003 1.73 ± 0.02 1.70 ± 0.02 1.71 ± 0.01 3.2 Scaling properties Despite using different values for the parameters, it is ob- served that the curves of the number of bacteria at the interface S versus the total number N show two power-law regimes. The first regime corresponds to initial compact structures S ∼ N 1/2, while the second regime corresponds to ramified structures with S ∼ N, as shown in Fig. 4a. These two behaviours are characteristic of the Eden and DLA models, respectively. To characterise the crossover between regimes, we compute the total number of bacteria N∗ at which the crossover takes place. To achieve this, we simply fit power- law functions in the tails of the S vs N curves and compute the intersection. More sofisticated methods for an auto- matic determination of crossovers can be found in [38]. After dividing N by N∗ in each of the data sets, the Y - axis is divided by some value S∗ looking for a satisfactory collapse of the curves. We found that the best collapse occurs when N∗ = S∗2, as shown in Fig. 4b. It can be seen that N∗ depends on the diffusion D and the concentration C, having an increasing relationship with both (Fig. 5). As part of the Springer Nature SharedIt initiative, a full-text view-only version of the paper can be accessed clicking here Lautaro Vassallo et al: On the growth of non-motile bacteria colonies 5 (a) (b) Fig. 4: Power-law regimes. (a) Results, averaging over all realizations, of the number of bacteria in the interface S versus the total number of cells N. (b) Collapsed curves of S versus N. Both plots are double-logarithmic. (a) (b) Fig. 5: Relationship between N∗ and the parameters. (a) N∗ is plotted against the nutrient concentration C, leaving the diffusion coefficient D constant; (b) the same but for D. (cid:40) (cid:40) (1) (3) (4) Taking these observations into account, an attempt is made to establish a scaling law. We know that the be- haviour of S is: S ∼ N 1/2, N (cid:28) N∗ N (cid:29) N∗ , N, where N∗ = N∗(C, D). The curves collapse dividing N and S by N∗ and S∗, respectively, so: S/S∗ ∼ (N/N∗)1/2, N/N∗ (cid:28) 1 N/N∗ (cid:29) 1 N/N∗, . (2) Then, having validated the relationship N∗ = S∗2 and proposing the scaling function (cid:40) f (x) ∼ const, x (cid:28) 1 x (cid:29) 1 x, , the relation between N and S can be written as: S = N 1/2f [( N N∗ )1/2]. This result suggests that if we allow N to grow suffi- ciently, branches will always be generated, after a critical number of cells is reached, dependent on the parameters C and D. 3.3 Multifractality of the growth probability The growth probability of each region of the colony can give information about why a certain pattern displays. Ev- ery cell duplicates when a nutrient particle is captured, so the growth probability is associated with the probability that a diffusing particle reaches the site where the cell is. We use two methods to estimate this probability, focus- ing on the final stage of the colony. The first one consists on counting how many nutrient particles are absorbed by each cell without letting it duplicate, i.e., the colony is "frozen" and the growth probability of each cell is com- puted dividing this counting by the total of particles in- corporated by the whole colony (we use approximately 105 particles). We will refer to this method as C.M. The disad- vantage with this method is that it does not estimate low probabilities well, because several million particles may be captured by the colony in total, but the internal re- gions may hardly incorporate any. Due to this, we also solve the Laplace equation ∇2φ = 0, where φ represents the nutrient concentration, by the relaxation method [39], where φ = 1 at infinity and φ = 0 along the perimeter of the colony, as it can be seen that the growth probability is proportional to the gradient of the potential ∇φ [29]. We use an iteration error of 10−5, after checking that the multifractal curves do not vary appreciably. In order to use this method, referred to as L.M. henceforth, properly, space has to be discretised, so some differences with the C.M. are expected. In Fig. 6, it is shown how uneven is the number of nutrient particles consumed between the outer and inner As part of the Springer Nature SharedIt initiative, a full-text view-only version of the paper can be accessed clicking here 6 Lautaro Vassallo et al: On the growth of non-motile bacteria colonies (a) Absorption > 1 (b) Absorption > 20 (c) Absorption > 80 Fig. 6: Example of the C.M. results for a branched colony. After 105 particles are captured, the green colour marks the cells who absorbed more than (a) 1 particle, (b) 20 particles and (c) 80 particles. regions of a ramified colony. This phenomenon is usually referred to as 'shadowing' or 'screening' effect and is more or less noticeable depending on the parameters. As the structures that emerge from the simulations are fractal, the proper way to study this effect is by the multifractal formalism, explained in the first section. In Fig. 7, the generalised dimension Dq>1 curve is plot- ted for different morphologies. It can be seen that the probability associated with a ramified colony presents a strong multifractality since Dq varies significantly with q. It is also included in Fig. 7 the curve for a diffusion-limited aggregate [40] for comparison. The standard deviations are presented in Table 3. Unfortunately, this analysis cannot be carried out in very compact colonies since the fractal regime is very short to be reliable or it is not observable. Nor can it be done in the early stages of the growth process for the same reason. In Fig. 8, the generalised dimension is plotted again, but now including the q < 1 interval. Only the results obtained by the L.M. can be used in this interval. The value of Dq(cid:28)−1 is very interesting, because it quantifies the shadowing effect, making evident the differences be- tween different morphologies. It is worth noting that Dq=0 should not be equal to the fractal dimension Df presented in the previous subsection because Dq=0 is the fractal di- mension of the support of the measure, i.e., the perimeter of the colony, while Df is the fractal dimension of the area. All of these characteristic points are summarised in Table 3, which also includes the values corresponding to diffusion-limited aggregation as a comparison [29,40]. The asymptotic values are estimated with q = 25 and q = −25. Note that the curves of the generalised dimension are always above the case of DLA. In the region of q > 0, taking into account the calculated standard deviations, (a) (b) Fig. 7: Generalised dimension curve for q > 1. Different icons correspond to different values of nutrient concentra- tion C and diffusion coefficient D. Red dashed line corre- sponds to DLA (colour online). Results using the (a) C.M. and (b) L.M. Fig. 8: Generalised dimension curve including negative values of q. Only the L.M. results are plotted. Different icons correspond to different values of nutrient concentra- tion C and diffusion coefficient D. Note that the curves do not collapse for q > 0, the difference between them cannot be seen on this scale (see Fig. 7). the differences are not as noticeable, but they are in the region of q < 0. In this region, where the measurement best distinguishes each case, they depart notoriously from the case of DLA. Always considering branched cases, it is observed that higher values are associated with higher C and D values, which can be understood if we associate this measure with the screening phenomenon. The larger are C and D, the thicker and narrower the branches and the fjords become, respectively, so the screening increases to the interior areas. Nevertheless, note that there is a As part of the Springer Nature SharedIt initiative, a full-text view-only version of the paper can be accessed clicking here Lautaro Vassallo et al: On the growth of non-motile bacteria colonies 7 Table 3: Some characteristic values of the generalised dimension Dq (the error is the standard deviation). DLA results found in [29, 40] are included for comparison. C = 1/20; D = 512 C.M. L.M. − C = 1/40; D = 512 C.M. L.M. − Dq=0 Dq=1 Dq=2 Dq(cid:29)1 Dq(cid:28)−1 1.08 ± 0.01 1.00 ± 0.02 0.82 ± 0.04 − 1.60 ± 0.01 1.08 ± 0.05 1.00 ± 0.09 0.75 ± 0.09 19.0 ± 1.5 1.09 ± 0.01 1.00 ± 0.02 0.82 ± 0.04 − 1.63 ± 0.01 1.10 ± 0.09 1.01 ± 0.12 0.79 ± 0.15 15.5 ± 1.5 C = 1/40; D = 1024 C.M. L.M. − 1.11 ± 0.01 1.04 ± 0.01 0.87 ± 0.04 − 1.53 ± 0.01 1.09 ± 0.08 1.01 ± 0.13 0.78 ± 0.15 18.7 ± 2.1 C = 1/80; D = 512 C.M. L.M. − Dq=0 Dq=1 Dq=2 Dq(cid:29)1 Dq(cid:28)−1 1.09 ± 0.01 1.00 ± 0.02 0.80 ± 0.03 − 1.64 ± 0.01 1.08 ± 0.07 0.98 ± 0.11 0.73 ± 0.14 14.0 ± 1.4 C = 1/80; D = 1024 C.M. L.M. − 1.13 ± 0.01 1.04 ± 0.01 0.85 ± 0.03 − 1.56 ± 0.01 1.08 ± 0.06 1.00 ± 0.10 0.76 ± 0.13 18.3 ± 1.9 DLA [29] 1.64 ± 0.01 1.04 ± 0.01 0.67 ± 0.03 − (cid:39) 9 [40] 1 0.92 0.66 − limit on how large the parameters can be. If C and D are too large, so the branches disappear, the screening effect is barely noticeable. The growth probability becomes almost uniform, and the multifractality should be lost. 4 Conclusions The goal of this work is the construction of a model based on basic theories of physics, capable of generating a va- riety of complex patterns observed in bacteria colonies. Under the hypothesis that there is a single collective move- ment mechanism behind the different morphologies (slid- ing), the different results are achieved by varying parame- ters of the environment outside the colony, without chang- ing the behaviour of the agents. Under these precepts, we manage to generate patterns from the most round and compact to extremely ramified, going through different intermediate morphologies. The different approaches used to characterise the structures also allow comparison with the two most studied models that predict patterns of bac- terial colonies, the Eden model and DLA. The fractal dimension analysis of simulations with branched colonies shows a fractal dimension compatible with experiments [10]. Although there are distinct dif- ferences between the present and the DLA model, their fractal dimensions are in good agreement, which suggests that the most ramified cases considered are close to the diffusion limit. On the other hand, the characterization using scal- ing laws show that there are two characteristic growth regimes, one compact and one branched. According to the relation found, the crossover between regimes occurs at a critical number of cells, that depends on the parameters for nutrition concentration C and nutrition diffusion D. Thus, branches will always be generated for finite values of these parameters. The multifractal measurement shows strong multifrac- tality for the ramified cases. In case the order q of the generalised dimension Dq is q > 0, the curves of Dq of the simulations of different values of C and D are close to each other, but in the region of q < 0, their differences become notoriously. All these curves are, however, always above the DLA curve and, as expected, approach to the DLA curve for lower values for C and D. Higher values are associated with higher C and D values because fjords into the interior are narrow. Unfortunately, a more in-depth comparison between simulation and experimental data cannot be carried out due to the lack of quantitative experimental data. To date, there are almost exclusively qualitative characterizations of the morphologies of bacterial colonies, which only in some rare cases provide additional information on the fractal dimension (which is not a complete indicator). Al- though the methods that we use in this work to calculate the generalised dimensions cannot be used in experiments, there are other methods that might be used, such as the one described by Ohta and Honjo [41], based on associ- ating probabilities according to the variation of the area occupied by the colony in a certain section. Only such a deeper experimental analysis would offer a complete char- acterization of the processes involved in the structure for- mation of bacterial colonies and would allow contrasting the proposed model with experiments. Conflict of interest Authors declare that there are no conflicts of interest. Acknowledgements L.A.B. and L.V. thanks UNMdP and CONICET (PIP 00443/2014) for financial support. D.H. thanks UNMdP and CONICET (PIP 0100629/2013) for financial support. As part of the Springer Nature SharedIt initiative, a full-text view-only version of the paper can be accessed clicking here 8 Lautaro Vassallo et al: On the growth of non-motile bacteria colonies 4. M.G. Mazza, Journal of Physics D: Applied Physics 49, 30. W. Kinsner, Journal of Information Technology Research 23. A. Marrocco, H. Henry, I.B. Holland, M. Plapp, S.J. Séror, B. Perthame, Mathematical Modelling of Natural Phenom- ena 5, 148 (2010) 24. E. Ben-Jacob, O. Schochet, A. Tenenbaum, I. Cohen, A. Czirók, T. Vicsek, Nature 368, 46 (1994) 25. F.D.C. Farrell, O. Hallatschek, D. Marenduzzo, B. Wa- claw, Physical Review Letters 111, 1 (2013) 26. B. Li, J. Wang, B. Wang, W. Liu, Z. Wu, Europhysics Letters (EPL) 30, 239 (1995) 27. P. Melke, P. Sahlin, A. Levchenko, H. Jönsson, PLoS Com- putational Biology 6, e1000819 (2010) 28. S.N. Santalla, S.C. Ferreira, Physical Review E 98, 022405 29. Y. Hayakawa, S. Sato, M. Matsushita, Physical Review A (2018) 36, 1963 (1987) (JITR) 1, 62 (2008) 31. A.L. Barabasi, H.E. Stanley, Fractal Concepts in Surface Growth (Cambridge University Press, Cambridge, 1995), ISBN 9780511599798, 0521483182 32. H. Hentschel, I. Procaccia, Physica D: Nonlinear Phenom- ena 8, 435 (1983) 33. T. Halsey, M. Jensen, L. Kadanoff, I. Procaccia, B. Shraiman, Physical review A 33, 1141 (1986) 34. A.B. Chhabra, C. Meneveau, R.V. Jensen, K.R. Sreeni- vasan, Physical Review A: Atomic, Molecular, and Optical Physics 40, 5284 (1989) 35. P.C. Ivanov, Q.D. Ma, R.P. Bartsch, J.M. Hausdorff, L.A. Nunes Amaral, V. Schulte-Frohlinde, H.E. Stanley, M. Yoneyama, Physical Review E - Statistical, Nonlinear, and Soft Matter Physics 79, 1 (2009) 36. H. Risken, The Fokker-Planck Equation, Vol. 18 of Springer Series in Synergetics (Springer Berlin Heidelberg, Berlin, Heidelberg, 1989), ISBN 978-3-540-61530-9 37. E. Catto, Iterative dynamics with temporal coherence, in Game developer conference (2005), Vol. 2, p. 5 38. A. Bashan, R. Bartsch, J.W. Kantelhardt, S. Havlin, Phys- ica A: Statistical Mechanics and its Applications 387, 5080 (2008) 39. J.M. Li, L. Lü, M.O. Lai, B. Ralph, in Image-based fractal description of microstructures (Springer US, Boston, MA, 2003), chap. 6, p. 129, ISBN 978-1-4419-5370-4 40. M. Matsushita, Y. Hayakawa, S. Sato, K. Honda, Physical Review Letters 59, 86 (1987) 41. S. Ohta, H. Honjo, Physical Review Letters 60, 611 (1988) Author contribution statement L.V. and D.H. designed the model. L.V. implemented the computational framework. All authors contributed to the analysis of the data and in the writing of the manuscript. References 1. G. De Magistris, D. Marenduzzo, Physica A: Statistical Mechanics and its Applications 418, 65 (2015) 2. J. Henrichsen, Bacteriological reviews 36, 478 (1972) 3. Mark J. McBride, Annual Review of Microbiology 55, 49 (2001) 203001 (2016) 5. J.W. Costerton, D.E. Calkwell, D.R. Kober, H. M.Lappin- Scott, Z. Lewandowski, D.E. Caldwell, D.R. Korber, H.M. Lappin-Scott, Annual Review of Microscopy 49, 711 (1995) 6. L. Hall-Stoodley, J.W. Costerton, P. Stoodley, Nature Re- views Microbiology 2, 95 (2004) 7. G. O'Toole, H.B. Kaplan, R. Kolter, Annual Review of Microbiology 54, 49 (2000) 8. H.C. Flemming, J. Wingender, Nature Reviews Microbiol- ogy 8, 623 (2010) 9. T. Matsuyama, M. Sogawa, Y. Nakagawa, FEMS microbi- ology letters 52, 243 (1989) 10. M. Matsushita, H. Fujikawa, Physica A: Statistical Me- chanics and its Applications 168, 498 (1990) 11. M. Matsushita, J. Wakita, H. Itoh, I. Ràfols, T. Mat- suyama, H. Sakaguchi, M. Mimura, Physica A: Statistical Mechanics and its Applications 249, 517 (1998) 12. M. Ohgiwari, M. Matsushita, T. Matsuyama, Journal of the Physical Society of Japan 61, 816 (1992) 13. H. Fujikawa, Physica A: Statistical Mechanics and its Ap- plications 189, 15 (1992) 14. K. Kawasaki, A. Mochizuki, M. Matsushita, T. Umeda, N. Shigesada, Journal of Theoretical Biology 188, 177 (1997) 15. M. Mimura, H. Sakaguchi, M. Matsushita, Physica A: Sta- tistical Mechanics and its Applications 282, 283 (2000) 16. C. Giverso, M. Verani, P. Ciarletta, Journal of the Royal Society, Interface / the Royal Society 12, 20141290 (2015) 17. I. Golding, Y. Kozlovsky, I. Cohen, E. Ben-Jacob, Phys- ica A: Statistical Mechanics and its Applications 260, 510 (1998) 18. M. Matsushita, J. Wakita, H. Itoh, K. Watanabe, T. Arai, Physica A: Statistical Mechanics and its Applications 274, 190 (1999) 19. Y. Kozlovsky, I. Cohen, I. Golding, E. Ben-Jacob, Physical Review E 59, 7025 (1999) 20. M. Matsushita, F. Hiramatsu, N. Kobayashi, T. Ozawa, Y. Yamazaki, T. Matsuyama, Biofilms 1, 305 (2004) 21. A.M. Lacasta, I.R. Cantalapiedra, C.E. Auguet, A. Peñaranda, L. Ramírez-Piscina, Physical review. E, Statistical physics, plasmas, fluids, and related interdisciplinary topics 59, 7036 (1999) 22. H. Tronnolone, A. Tam, Z. Szenczi, J.E. Green, S. Bal- asuriya, E.L. Tek, J.M. Gardner, J.F. Sundstrom, V. Ji- ranek, S.G. Oliver et al., Scientific Reports 8, 1 (2018)
1807.09124
2
1807
2018-09-10T03:28:34
Flapping-pattern change in small and very small insects
[ "physics.bio-ph", "physics.flu-dyn" ]
Medium and large insects in normal hovering have horizontal, planar up- and downstrokes1-4. The lift of the two half-strokes, generated by the leading-edge vortex, provides the weight-supporting vertical force. But for small insects (wing length R less than about 4 mm and Reynolds number Re very low, about 80 to 10), because of the large effect of air viscosity (as Re becomes very low, moving in air is like in oil), sufficient vertical force could not be produced if using the above wing kinematics. Small insects must use different flapping mode. Here, through analyzing flight data from our recent studies on a relatively-large small insect (fruitfly: R=3 mm, Re=80) and a very small insect (wasp: R=0.5 mm, Re=10), we put forward a hypothesis on how the flapping pattern will change: as insect-size or Re decreasing, a deeper and deeper U-shape upstroke will be used to overcome the viscous effect. And we test this hypothesis by measuring the wing kinematics for species of different sizes to obtain data for Re ranging from 80 to 10 and by computing the aerodynamic forces. The data and computation support our hypothesis: the planar upstroke changes to U-shape upstroke which becomes deeper as size or Re becomes smaller; for relatively-large small insects, the U-shape upstroke produces a larger vertical force than a planar upstroke by having a larger wing velocity, and for very small insects, the deep U-shape upstroke produces a large transient drag that points almost upwards by fast downward acceleration of the wing, providing the required vertical force.
physics.bio-ph
physics
Flapping-pattern change in small and very small insects Yu Zhu Lyu, Hao Jie Zhu and Mao Sun* Institute of Fluid Mechanics, Beihang University, Beijing, China (* [email protected]) Medium and large insects in normal hovering have horizontal, planar up- and downstrokes1-4. The lift of the two half-strokes, generated by the leading-edge vortex5-10, provides the weight-supporting vertical force. But for small insects (wing length R less than about 4 mm and Reynolds number Re very low, about 80 to 10)11, because of the large effect of air viscosity (as Re becomes very low, moving in air is like in oil), sufficient vertical force could not be produced if using the above wing kinematics12,13. Small insects must use different flapping mode. Here, through analyzing flight data from our recent studies on a relatively-large small insect (fruitfly: R≈3 mm, Re≈80)14 and a very small insect (wasp: R≈0.5 mm, Re≈10)15, we put forward a hypothesis on how the flapping pattern will change: as insect-size or Re decreasing, a deeper and deeper U-shape upstroke will be used to overcome the viscous effect. And we test this hypothesis by measuring the wing kinematics for species of different sizes to obtain data for Re ranging from 80 to 10 and by computing the aerodynamic forces. The data and computation support our hypothesis: the planar upstroke changes to U-shape upstroke which becomes deeper as size or Re becomes smaller; for relatively-large small insects, the U-shape upstroke produces a larger vertical force than a planar upstroke by having a larger wing velocity, and for very small insects, the deep U-shape upstroke produces a large transient drag that points almost upwards by fast downward acceleration of the wing, providing the required vertical force. About half of the existing winged insect species are of small size (wing-length R≈0.5-4 mm)11. But much of what we know about the biomechanical mechanisms of insect flight is derived from studies on medium and large insects (R≈5-50 mm) insects5-10, such as flies, honey-bees and hawkmoths. The non-dimensional parameter 1 representing the effect of air viscosity is the Reynolds number (Re): lower Re means the wing moving in a more viscous flow. For the flapping wing of an insect, Re is approximately proportional to the square of R (Re is defined using the mean chord-length cm and the mean wing speed U=2Φfr2 where Φ is the stroke aptitude, f the stroke frequency and r2 the radius of gyration of wing). Thus the wings of the small insects operate at very low Re, on the order of 80-10. At this range of Re, moving in the air is like in oil. Medium and large insects in normal hovering beat their wings approximately in a horizontal plane (Fig. 1a)1-4 and the wings operate at Reynolds number (Re) about 100-3500. During the downstroke or upstroke, a lift, and a drag that is a little smaller, are produced (Fig. 1a). The lift provides the weight supporting vertical force; the drag in the downstroke cancels out that in the upstroke and the flapping-cycle mean horizontal force is zero. The aerodynamic forces are generated mainly by the leading-edge vortex (LEV) that attaches to the wing in the entire up- or downstroke, which is referred to as the delayed-stall mechanism5-10. However, if the small insects flap their wings as their larger counterparts, sufficient aerodynamic force cannot be produced because of the very strong viscous effects: the LEV is significantly defused and little lift can be generated, while the drag is very large12,13. The small insects must have used different wing kinematics and aerodynamic mechanisms from those of the medium and large insects. 2 Figure 1. (a) Left: the wing-tip trajectory (projected onto the symmetrical plane) of a dronefly Eristalis tenax (ET)4, a relatively large insect; right: the motion of a section of the wing. (b) Left: reference frame and Euler angles defining the wing kinematics: (x, y, z) are coordinates in a system with its origin at the wing root and with x-axis points horizontally backwards and z-axis points vertically upwards and y-axis points to the leftt of the insect,  is the positional angle (in the stroke plane), ψ the pitch angle,  deviation angle, and  the stroke-plane angle; right: measured Euler angles of biting-midge FG1 (mean±s.d.; n=30 wingbeats), biting-midge DF1 (mean±s.d.; n=30 wingbeats), gall midge AS1 (mean±s.d.; n=6 wingbeats) and thrip FO1 (mean±s.d.; n=6 wingbeats). T, stroke period. From our recent studies on hovering of a relatively-large small insects, fruitfly Drosophila virilis (DV) (R≈3 mm, Re≈80)14 and a very small insect, wasp Encarsia Formosa (EF) (R≈0.5 mm, Re≈10)15, we observe that the wasp has a very deep U-shape upstroke and the fruitfly also has a U-shape upstroke but much shallower, and that the deep U-shape upstroke can generate a large transient-drag that pointed almost upwards, enhancing the vertical force production. Based on this observation, we put forth a hypothesis that as insect-size or Re decreasing, deeper and deeper U-shape upstroke would be used to overcome the viscous effect. We test this 3 hypothesis by measuring the wing kinematics of more species of different sizes and obtained data for Re ranging from about 80 to 10. A few years earlier, we measured the wing motion of vegetable leafminers Liriomyza sativae (LS) (R≈1.5 mm and Re≈40)16. Here the wing kinematics of four more species are measured: biting-midges Forcipomia gloriose (FG) and Dasyhelea flaviventris (DF), gall-midge Anbremia sp. (AS) and Thrip Frankliniella occidentalis (FO). Their Re is 30, 24, 17 and 14, respectively, and is between those of LS (40) and EF (10). Fig. 1b shows the measured Euler angles of wing in four individuals, each from one of the four species (Supplementary Videos 1-4; the videos can also be seen at: msun.buaa.edu.cn). For each of the four species, data of another four individuals (Supplementary Table 1) were also measured; within a species, the results are similar (Supplementary Fig. 1). Using data in Fig. 1b, stroke diagrams showing the flapping mode of the four insects are plotted in Fig. 2 (those of small insects DV14, LS16 and EF15 and a large insect, dronefly Eristalis tenax (ET)4, are also included). The results support our hypothesis: as size or Re decreasing, the insect has a deeper and deeper U-shape upstroke (Fig. 2b-h). For the two very small insects, thrip FO and wasp EF, the downstroke is also U-shaped, but a much shallower one (Fig. 2g, h), which is the result of the 'fling' motion discovered by Weis-Fogh1 (described in detail elsewhere15,17-20). Within each of the species, all the individuals have the same flapping pattern (Supplementary Fig. 2). 4 5 Figure 2. (a) through (h): stroke diagrams show the wing motions of the insects; the solid curve indicates the wing-tip trajectory (projected onto the x-y plane); black lines indicate the orientation of the wing at various times in one stroke cycle, with dots marking the leading edge; a black dot defines the wing-root location on the insect body; the blue arrow represents the velocity of the wing at the radius of gyration; the red arrow represents the total aerodynamic force of the wing. (a*) through (h*): vertical forces in one stroke cycle (black line); contribution by the lift (blue line) and that by drag (orange line); gray background indicates the upstroke. In the early part of the U-shape upstroke (t/T≈0-0.2), the wing accelerates downwards and backwards with wing surface almost horizontal and at a rather large angle of attack; the smaller the insect or the lower the Re, the larger the acceleration (see the change in wing speed in Fig. 2b-h; here, the velocity at the radius of gyration of wing is used to represent the velocity of the wing). In the later part of the U-shape upstroke, generally the wing moves slower (Fig. 2e-h); as Re becomes very small, the wing moves almost vertically upwards with the wing surface vertical and the angle of attack close to zero (Fig. 2e-h). To assess how the weight-supporting force is generated, the flow and forces on the wings were computed using an experiment-data validated flow solver. Fig. 2a*-h* give the computed vertical forces (FV) for the eight insects shown in Fig. 2a-h (the corresponding horizontal forces are given in Supplementary Fig. 3; the forces for the other individuals of each species given in Supplementary Fig. 4). For the large insect ET (Fig. 2a*), FV in the upstroke (t/T=0-0.5) is similar to that in the downstroke (t/T=0.5-1). But for the small insects (Fig. 2b*-h*), the U-shape upstroke produces a large FV peak, while the planar downstroke produces a relatively small FV. In general, as size or Re deceasing, the FV peak in the U-shape upstroke becomes higher and narrower. For the two very small insects (FO and EF), a smaller FV peak is also 6 produced at the early downstroke (Fig. 2g* and h*, t/T=0.55-0.7) by the 'fling' motion aforementioned. The vertical and horizontal forces of a wing come from its lift and drag, which are the components of the total aerodynamic force that are perpendicular and parallel to the wing-velocity, respectively. For the large insect ET (Fig. 2a*), almost all the vertical force is contributed by the lift. For the relatively large small insects DV and LS (Fig. 2b* and c*), vertical force is mainly (about 90%) is contributed by the lift. As insects become smaller (Fig. 2d*-h*), the drag has more and more contribution; for the two smallest insects, FO and EF, about 70% vertical-force is from the drag. Figuere. 3. Non-dimensional spanwise vorticity contours in the section at the radiius of gyration, at verious times during the U-shape upstroke. Vorticity is nondimensionalized by U/cm; solid and dashed lines denote the anticlockwise and clockwise vorticity, respectively. The magnitude of the non-dimensional vorticity at the outercontour is 2 and the contour interval is 1. To explain the large force in the U-shape upstroke, vorticity-fields are plotted for 7 the period in which the large aerodynamic force is produced. We first consider DV and LS, whose size is relatively large and Re relatively high (about 80-40). Fig. 3b shows the vorticity plots of LS1 in the period of t/T=0.18-0.34: a LEV attaches and moves with the wing, indicating that the force is produced by the delayed-stall mechanism. It can be shown that the planar downstroke also use the delayed-stall mechanism to produce the force. The reason for the U-shaped upstroke producing a larger force than the planar downstroke is that it has a larger wing velocity (see Fig. 2c). The forces of DV can be similarly explained. Next we consider the smaller insects (Figs. 3c-e). As an example, we look at the vorticity plots of FO (Fig. 3e): during the very short period (t/T=0.14-0.22), counter clockwise vorticity is continuously produced around the leading edge of the wing and clockwise vorticity around the trailing edge. This would result in a large time rate of change in the first moment of vorticity, giving the large aerodynamic force21. This force producing mechanism is called as 'rowing mechanism'15,22: the wing accelerates fast from zero velocity at a very high angle of attack, producing a large transient drag (here the drag points almost upwards, giving the large vertical force). The same is true for FG, DF, AS and EF. 8 Figuere. 4. (a) through (h): vertical force produced if the upstroke was planar (green dashed line), compared with that using the real wing kinematics which has U-shape upstroke (black line). To show the advantage of having a U-shape upstroke for the small insects, we made test calculations in which both the down- and upstrokes were planar and horizontal, like those of the larger insects, while Re being kept the same. The computed vertical forces for the eight insects are shown in Fig. 4a-h (the corresponding horizontal forces are given in Supplementary Fig. 5). It is seen that the planar upstroke produce much less vertical force than the U-shape upstroke and it is more so as Re becomes smaller. For the fruitfly DV (Re≈77), the mean vertical force produced by the U-shape upstroke is 1.43 times of that by the planar upstroke. For LS (Re≈40), FG (Re≈30), DF (Re≈24), AS (Re≈17), FO (Re≈14) and EF (Re≈10), the corresponding numbers are 1.82, 1.75, 1.92, 2.27, 2.27, and 4.17, respectively. This 9 shows that the small insects use the novel motion, the U-shape upstroke, to overcome the problem of insufficient lift at very small Reynolds number, encountered by the commonly used flapping kinematics. Taken together, our findings show that the small insects change their flapping mode to solve the low-Re problem: The planar upstroke changes to U-shape upstroke, and as size or Re becomes smaller, a deeper U-shape upstroke is employed; for relatively-large small insects (Re about 80-40), the U-shape upstroke produces a larger vertical force than a planar upstroke by having a larger wing velocity, and for very small insects (Re below about 40), the deep U-shape upstroke produces a large transient drag that points almost upwards by fast downward acceleration of the wing, providing the required vertical force. Methods Insects. Thrips (FO) were acquired from the Laboratory of Institute of Vegetables and Flowers, Chinese Academy of Agricultural Sciences, which were descendents of wild-caught Frankliniella occidentalis. Biting-midges Forcipomia gloriose (FG) and Dasyhelea flaviventris (DF), gall midges Anbremia sp. (AS) were netted in a suburb of Beijing in June to August 2017. High-speed filming. The near-hover flights of the small insects in transparent flight chambers were filmed using three orthogonally aligned synchronized high-speed cameras (FASTCAM Mini UX100, Photron Inc., San Diego, CA, USA) mounted on an optical table (Supplementary Fig. 6a). The size of the flight chamber is 50×50×50 10 mm3 for biting-midge Forcipomia gloriose (FG) and 34×34×34 mm3 for biting-midge Dasyhelea flaviventris (DF), gall-midge Anbremia sp. (AS) and Thrip Frankliniella occidentalis (FO). Each camera was equipped with a 60 mm micro-Nikkor lens and 12 mm extension tube. For FG, the cameras were set to 10,000 (or 10500) frames per second (resolution 1280×496 or 1280×472) pixels and shutter speed to 20 s; for DF, AS and FO, to 8,000 frames per second (resolution 1280×624 pixels) and shutter speed to 20 s. Each camera view was backlit using a 50 W integrated red light emitting diode (LED; luminous flux, 4000 lm; wavelength, 632 nm) and two lenses were used to make the light uniform. The synchronized cameras were manually triggered when the insect was observed to fly steadily in the filming area (approximately 6×6×6 mm3) which represented the intersecting field of views of the three cameras. The experiment was performed at temperature 25-27C and relative humidity 50-60%. Kinematics reconstruction. The orthogonally aligned cameras were calibrated by using a flat glass panel with a high accuracy black-and-white checkerboard pattern printed on it. The calibration gave the intrinsic and extrinsic parameters of each camera which determined the transform matrix of the camera4,23. The method used to extract the 3D body and wing kinematics from the filmed data was developed in previous works of our group4,23,15. The body and wings were represented by models (Supplementary Fig. 6b): the model of the body was two lines perpendicular to each other, which were the line connecting the head and the end of the abdomen and the line connecting the two wing hinges (Supplementary Fig. 6b); 11 the model of a wing was the outline of the wing obtained by scanning the cut-off wing (Supplementary Fig. 6d) and the wing model can have a spanwise bending, represented by the maximum bending displacement (Supplementary Fig. 6c). An interactive graphic user interface developed using MATLAB (v.7.1, The Mathworks, Inc., Natick, MA, USA) was used to determine the positions and orientations of the body and the wings: the positions and orientations of the models of the body and wings were adjusted until the best overlap between a model image and the displayed frame was achieved in three views, and at this point the positions and orientations of these models were taken as the positions and orientations of the body and the wings. The fitting process was manually done. More detailed description of the method can be found elsewhere4,23,15. This process gives the position and orientation of the body, the wing root positions, the Euler angles and the maximum bending displacement of the wings. We processed 360 wingbeats in total, over 20 flight sequences from 20 individuals: 30 wingbeats for each of the 5 biting-midges Forcipomia gloriose and for each of the 5 biting-midges Dasyhelea flaviventris (these two species have high wingbeat frequency); and 6 wingbeats for each of the 5 gall-midge Anbremia sp. and for each of the 5 thrip Frankliniella occidentalis. Aerodynamic-force computation. The flows around and the aerodynamic force acting on the insects were computed using the method of computational fluid dynamics (CFD). For medium and large insects at hovering flight, it had been shown that aerodynamic interaction between the body and the wings was negligibly small: 12 the aerodynamic force in the case with body/wing interaction was less than 2.5% different from that without body/wing interaction24. Our computations showed that this also true for the small insects. Therefore, in the present CFD model, only the two wings were considered. The planform of a model wing is approximately the same as that of the corresponding insect wing (the wing planforms for FG, DF, AS and FO are shown in Supplementary Fig. 6d; those for ET, DV, LS and EF in our previous works4,14,16,15); the section of the model wing is a flat plat of 3% thickness with rounded leading and trailing edges. The incompressible Navier -- Stokes equations were solved over moving overset grids because there are relative movements between the left and right wings. There was a body-fitted curvilinear grid for each of the wings and a background Cartesian grid which extends to the far-field boundary of the domain (Supplementary Fig. 7a). The flow solver, which was based on an artificial compressibility method developed by Rogers et al.25, was the same as that used in several previous studies of our group8,15,19; its detailed description can be found there. The solver has been validated by comprehensive tests15: comparison with measured data of a revolving wing (Re=500)26 and of a flapping wing (Re=180)27; comparison with theoretical (Stokes or Oseen solutions) and measured results for a sphere at very low Re (5-20)28. Grid resolution tests were conducted to ensure that the flow calculations were grid independent. For the dronefly (ET) wing (Re=720), three grid-systems were considered. For grid-system 1, the wing grid had dimensions 41 61 43 in the 13 normal direction, around the wing, and in the span-wise direction, respectively (first layer grid thickness was 0.0015c); the body grid had dimensions 66 51 35 along the body, in the azimuthal direction and in the normal direction, respectively; the background grid had dimensions 81  81  81 in the X, Y and Z directions, respectively. For grid-system 2, the corresponding grid dimensions were 61 91 65, 99 77 53 and 121 121 121 (0.001c). For grid-system 3, the corresponding grid dimensions were 91 135 96, 149 118 80and 181 181 181 (0.00067c). For all the three grid-systems, grid points of the background grid concentrated in the near field of the wings where its grid density was approximately the same as that of the outer part of the body-grid. Three grid-systems similar to the above were also used for the small wasp (EF) wing (Re=10). The aerodynamic forces computed are shown in Supplementary Fig. 7b and c. It is seen that there is almost no difference between the force coefficients calculated by the three grid-systems. Calculations were also conducted using a larger computational domain. The domain was enlarged by adding more grid points to the outside of the background grid of grid-system 2. The calculated results showed that there was no need to put the outer boundary further than that of grid-system 2. The above results showed that grid-system 2 was proper for the wings of ET and EF. The wings of fruitfly DV, biting midge DF and gall midge AS had similar aspect ratio as that of ET or EF and operated at Re between those of ET and EF, therefore grid-system 2 was also used for insects of these 4 species. For biting midge FG, whose wing had a much larger aspect ratio, the wing grid in grid-system 2 was changed to 65×91×70 (more points 14 in the span-wise direction); and for thrip FO, whose wing had a smaller aspect ratio, the wing grid was changed to 65×91×56 (less points in the span-wise direction). The effect of time step value was also studied and it was found that a numerical solution effectively independent of the time step was achieved if the time step value was ≤ T/440, and this value was used in all the calculations. References 1. Weis-Fogh, T. Quick estimates of flight fitness in hovering animals, including novel mechanisms for lift production. J. Exp. Biol. 59, 169-230 (1973). 2. Ellington, C. P. The aerodynamics of hovering insect flight. III. Kinematics. Phil. Trans. R. Soc. Lond. B 305, 41-78 (1984). 3. Walker, S. M., Thomas, A. L. R. & Taylor, G. K. Deformable wing kinematics in free-flying hoverflies. J. R. Soc. Interface 7, 131-142 (2010). 4. Liu, Y. P. & Sun, M. Wing kinematics measurement and aerodynamics of hovering droneflies. J. Exp. Biol. 211, 2014-2025 (2008). 5. Ellington, C. P., Van den Berg, C., Willmott, A. P. & Thomas, A. L. R. Leading-edge vortices in insect flight. Nature 384, 626-630 (1996). 6. Dickinson, M. H., Lehmann, F. O. & Sane, S. P. Wing rotation and the aerodynamic basis of insect flight. Science 284, 1954-1960 (1999). 7. Bomphrey, R. J., Srygley, R. B., Taylor, G. K. & Thomas, A. L. R. Visualizing the Flow around Insect Wings. Phys. Fluids 14, S4-S4 (2002). 8. Sun, M. & Tang, J. Unsteady aerodynamic force generation by a model fruit fly 15 wing in flapping motion. J. Exp. Biol. 205, 55-70 (2002). 9. Chin, D. D. & Lentink, D. Flapping wing aerodynamics: from insects to vertebrates. J. Exp. Biol. 219, 920-932 (2016). 10. Lee, J., Choi, H. & Kim, H. Y. A scaling law for the lift of hovering insects. J. Fluid Mech. 782, 479-490 (2015). 11. Dudley, R. The biomechanics of insect flight: form, function, evolution. (Princeton University Press, 2002). 12. Wu, J. H. & Sun, M. Unsteady aerodynamic forces of a flapping wing. J. Exp. Biol. 207, 1137-1150 (2004). 13. Miller, L. A. & Peskin, C. S. When vortices stick: an aerodynamic transition in tiny insect flight. J. Exp. Biol. 207, 3073-3088 (2004). 14. Meng, X. G. & Sun, M. Aerodynamics and vortical structures in hovering fruitflies. Phys. Fluids 27, 031901 (2015). 15. Cheng, X. & Sun, M. Very small insects use novel wing flapping and drag principle to generate the weight-supporting vertical force. J. Fluid Mech. 2018 (in press). Also see: preprint at http://arXiv.org/bio-ph/1807.05629 (2018). 16. Cheng, X. & Sun, M. Wing-kinematics measurement and aerodynamics in a small insect in hovering flight. Sci. Rep. 6, 25706 (2016). 17. Spedding, G. R. & Maxworthy, T. The generation of circulation and lift in a rigid two-dimensional fling. J. Fluid Mech. 165, 247-272 (1986). 18. Miller, L. A. & Peskin, C. S. A computational fluid dynamics of 'clap and fling' in the smallest insects. J. Exp. Biol. 208, 195-212 (2005). 16 19. Sun, M. & Yu, X. Aerodynamic force generation in hovering flight in a tiny insect. AIAA J. 44, 1532-1540 (2006). 20. Arora, N., Gupta, A., Sanghi, S., Aono, H. & Shyy, W. Lift-drag and flow structures associated with the "clap and fling" motion. Phys. Fluids 26, 071906 (2014). 21. Wu, J. C. Theory for Aerodynamic Force and Moment in Viscous Flows. AIAA J. 19, 432-441 (1981). 22. Zhu, H. J. & Sun, M. Unsteady aerodynamic force mechanisms of a hoverfly hovering with a short stroke-amplitude. Phys. Fluids 29, 081901 (2017). 23. Mou, X. L., Liu, Y. P. & Sun, M. Wing motion measurement and aerodynamics of hovering true hoverflies. J. Exp. Biol. 214, 2832-2844 (2011). 24. Aono, H., Liang, F. & Liu, H. Near- and far-field aerodynamics in insect hovering flight: an integrated computational study. J. Exp. Biol. 211, 239-257 (2008). 25. Rogers, S. E., Kwak, D. & Kiris, C. Steady and unsteady solutions of the incompressible Navier-Stokes equations. AIAA J. 29, 603 -- 610 (1991). 26. Manar, F. & Jones, A. R. The Effect of Tip Clearance on Low Reynolds Number Rotating Wings. 52nd Aerosp. Sci. Meet. 1452 (2014). 27. Han, J.-S., Chang, J. W. & Han, J.-H. The advance ratio effect on the lift augmentations of an insect-like flapping wing in forward flight. J. Fluid Mech. 808, 485 -- 510 (2016). 28. White, F. M. Viscous fluid flow, Vol. 2, Boston: McGraw-Hill Higher Education. 1991. 17 Supplementary Information Supplementary Videos 1-4: near-hover flight of FG1, DF1, AS1 and FO1, respectively. In each video, the left, middle and right parts of the movie show the flight captured by the top-view camera and two side-view cameras, respectively. For FG1, Playback speed is 10fps, approximately 0.1% of the actual speed of the movie; for DF1, 8fps, approximately 0.1%; for AS1 and FO1, 24fps, approximately 0.3%. Acknowledgements This research was supported by grants from the National Natural Science Foundation of China (11832004, 11721202). Author Contributions M.S. conceived the experimental design; Y.Z.L. and H. J. Z. designed and constructed the apparatus and all authors contributed to data collection; Y.Z.L. processed the raw data to extract detailed kinematics and performed the CFD computations; M.S drafted the manuscript; all authors contributed to data interpretation and manuscript preparation. Author Information The authors declare no competing financial interests. Correspondence and requests for materials should be addressed to M.S. ([email protected]) 18 Supplementary Figure 1. Wing kinematics for each of the individuals. Measured Euler angles of wing and the maximum span-wise bending displacement ε (dm/R) in one stroke cycle. (a) Biting midge (FG). (b) Biting midge (DF). (c) Gall midge Anbremia sp. (AS). (d) thrip Frankliniella occidentalis (FO). 19 Supplementary Figure 2. Stroke diagrams show the wing motions of the insects. (a) through (d): biting-midge FG, biting-midge DF, gall midge Anbremia sp. and thrips Frankliniella occidentalis. Solid curve indicates the wing-tip trajectory (projected onto the x-y plane); black lines indicate the orientation of the wing at various times in one stroke cycle, with dots marking the leading edge; black dot defines the wing-root location on the insect body; blue arrow, velocity of the wing at the radius of gyration; red arrow, total aerodynamic force of the wing. (Similar plots can be obtained for ET4, DV14, LS16 and EF15). Supplementary Figure 3. Horizontal forces (black line) for the eight insects shown in Fig. 2a-h; contributions by the lift (blue line) and that by drag (orange line). The corresponding vertical forces are in Fig. 2 a*-h*. 20 Supplementary Figure 4. Vertical and horizontal forces for the other four individuals of each species. 21 Supplementary Figure 5. Horizontal force produced if the upstroke was planar (green dashed line), compared with that using the real wing kinematics which has U-shape upstroke (black line). Supplementary Figure 6. Experimental setup, kinematics extraction and wing bending definition. (a) A sketch showing the flight chamber and cameras. Each camera view is backlit using a integrated red light emitting diode (LED). (b) Extraction of body and wing kinematics. (c) Definition of the spanwise bending of wing. dm, maximum bending displacement, assumed to be at 0.4R. (d) Outline of the wings. 22 Supplementary Figure 7. (a) Portions of a computational grid-system. (b and c) Time courses of the vertical and horizontal forces in one cycle, calculated with three grid-systems, for ET and EF respectively. 23 Supplementary Table 1. Flight parameters. ID FG1 FG2 FG3 FG4 FG5 DF1 DF2 DF3 DF4 DF5 AS1 AS2 AS3 AS4 AS5 Re 29.8 35.9 34.8 31.9 29.3 23.7 23.9 19.2 20.7 21.0 17.4 14.9 18.8 20.7 17.3 R S (mm) (mm2) 1.30 1.40 1.40 1.34 1.38 0.92 0.95 0.88 0.84 0.92 1.33 1.28 1.36 1.47 1.39 0.38 0.44 0.46 0.41 0.45 0.25 0.22 0.20 0.20 0.23 0.51 0.43 0.57 0.65 0.54 r2/R 0.60 0.60 0.60 0.60 0.59 0.60 0.63 0.62 0.61 0.62 0.63 0.64 0.63 0.62 0.63 lr (mm) lb (mm) 0.53 0.57 0.59 0.55 0.58 0.27 0.28 0.25 0.28 0.29 0.25 0.20 0.25 0.31 0.27 2.29 2.36 2.47 2.33 2.40 1.34 1.21 1.16 1.14 1.28 1.30 1.05 1.12 1.38 1.22 f (Hz) 1117 1112 1019 1107 1016 723 781 724 739 698 238 230 230 226 213 Ф (º) 52 55 55 52 49 98 97 96 103 96 102 106 103 102 108  (º) 5.3 2.6 -4.1 7.2 12.6 0.9 -0.2 7.7 7.8 5.8 4.8 1.3 1.9 -3.4 -1.2 u (m/s) w (m/s) 0.00 0.02 0.10 0.12 0.00 0.01 0.00 0.00 0.05 0.16 0.14 0.09 0.10 0.11 0.16 0.00 0.14 0.21 0.07 0.00 0.01 0.14 0.00 0.02 0.05 0.08 0.12 0.09 0.04 0.03 FV, m (N) 1.849 2.350 2.622 2.216 1.679 0.840 1.056 0.690 0.690 0.717 0.463 0.340 0.481 0.724 0.634  (º) 4.5 1.3 -1.4 0.9 -8.6 -6.6 -3.3 -6.5 -7.0 -4.4 -7.7 -9.5 -9.3 -3.5 -7.6 239 223 213 225 231 140 137 143 141 134 6.3 18.3 9.9 -4.4 14.7 0.13 0.22 0.12 0.14 0.22 4.0 2.6 0.9 14.9 0.4 1.13 1.23 1.09 1.18 1.27 0.30 0.29 0.33 0.31 0.34 13.5 11.9 13.4 12.6 14.3 0.76 0.79 0.83 0.79 0.86 0.59 0.59 0.59 0.57 0.61 0.28 0.27 0.26 0.26 0.29 FO1 FO2 FO3 FO4 FO5 Re, Reynolds number; R, wing length; S, area of wing; r2, radius of gyration of wing; lr, the distance between the left and right wing-roots; lb, body length. f and Ф, stroke frequency and amplitude, respectively; , stroke plane angle; u and w, horizontal and vertical velocities of body, respectively; FV,m, mean vertical force; , angle from the vertical of the mean force vector. 0.167 0.232 0.223 0.205 0.206 0.17 0.00 0.08 0.07 0.11 24
1609.00235
1
1609
2016-09-01T13:45:56
Electronic structure and carrier transfer in B-DNA monomer polymers and dimer polymers: Stationary and time-dependent aspects of wire model vs. extended ladder model
[ "physics.bio-ph" ]
We employ two Tight-Binding (TB) approaches to study the electronic structure and hole or electron transfer in B-DNA monomer polymers and dimer polymers made up of $N$ monomers (base pairs): (I) at the base-pair level, using the on-site energies of base pairs and the hopping integrals between successive base pairs, i.e., a wire model and (II) at the single-base level, using the on-site energies of the bases and the hopping integrals between neighboring bases, i.e., an \textit{extended} ladder model since we also include diagonal hoppings. We solve a system of $MD$ ("matrix dimension") coupled equations [(I) $MD$ = $N$, (II) $MD$ = $2N$] for the time-independent problem, and a system of $MD$ coupled $1^\text{st}$ order differential equations for the time-dependent problem. We study the HOMO and the LUMO eigenspectra, the occupation probabilities, the Density of States (DOS) and the HOMO-LUMO gap as well as the mean over time probabilities to find the carrier at each site [(I) base pair or (II) base)], the Fourier spectra, which reflect the frequency content of charge transfer (CT) and the pure mean transfer rates from a certain site to another. The two TB approaches give coherent, complementary aspects of electronic properties and charge transfer in B-DNA monomer polymers and dimer polymers.
physics.bio-ph
physics
Electronic structure and carrier transfer in B-DNA monomer polymers and dimer polymers: Stationary and time-dependent aspects of wire model vs. extended ladder model K. Lambropoulos,1 M. Chatzieleftheriou,1, ∗ A. Morphis,1 K. Kaklamanis,1 R.Lopp,1, † M. Theodorakou,1 M. Tassi,1 and C. Simserides1, ‡ 1National and Kapodistrian University of Athens, Faculty of Physics, Panepistimiopolis, 15784 Zografos, Athens, Greece (Dated: October 2, 2018) We employ two Tight-Binding (TB) approaches to study the electronic structure and hole or electron transfer in B-DNA monomer polymers and dimer polymers made up of N monomers (base pairs): (I) at the base-pair level, using the on-site energies of base pairs and the hopping integrals between successive base pairs, i.e., a wire model and (II) at the single-base level, using the on-site energies of the bases and the hopping integrals between neighboring bases, i.e., an extended ladder model since we also include diagonal hoppings. We solve a system of M D ("matrix dimension") coupled equations [(I) M D = N , (II) M D = 2N ] for the time-independent problem, and a system of M D coupled 1st order differential equations for the time-dependent problem. We study the HOMO and the LUMO eigenspectra, the occupation probabilities, the Density of States (DOS) and the HOMO-LUMO gap as well as the mean over time probabilities to find the carrier at each site [(I) base pair or (II) base)], the Fourier spectra, which reflect the frequency content of charge transfer (CT) and the pure mean transfer rates from a certain site to another. The two TB approaches give coherent, complementary aspects of electronic properties and charge transfer in B-DNA monomer polymers and dimer polymers. PACS numbers: 87.14.gk, 82.39.Jn, 73.63.-b I. INTRODUCTION Today, remarkable parts of the physical, chemical, bio- logical and medical communities as well as a broad spec- trum of other scientists and engineers is interested in charge transfer (CT) in biological systems. CT is the ba- sis of many biological processes e.g. in various proteins1 including metalloproteins2 and enzymes3, with medical and bioengineering applications4,5. CT plays a central role in DNA damage and repair6 -- 8. CT might also be an indicator to discriminate between pathogenic and non- pathogenic mutations at an early stage9. DNA plays a key role in the development, function and reproduction of living organisms, because the sequence of its bases (adenine, guanine, thymine, cytosine) car- ries their genetic code, hence, its study is usually associ- ated with molecular biology and genetics. However, its remarkable properties have spurred in recent years the interest of a broad interdisciplinary community. From the perspective of physics, its electronic structure and its CT properties are studied with the aim of understanding its biological functions as well as its potential applica- tions in nanotechnology (e.g. nanosensors, nanocircuits, molecular wire). At least for twenty years, we try to understand car- rier movement through DNA10 -- 16. Today, we know that many external factors related to the environment, like aqueousness and presence of counterions, extraction pro- cess, conduct quality with electrodes, purity, substrate and so on, influence carrier movement. This leads to the need of a deeper understanding of endogenous fac- tors affecting carrier movement in DNA, like base-pair sequence and geometry. Maybe the most important en- dogenous factor is the base-pair sequence, to which this article is devoted, too. Additionally, we have to discriminate between the words transport (usually implying the use of electrodes), transfer, and migration (a transfer over rather long dis- tances). The carriers (electrons or holes) can be either inserted via electrodes or generated by UV irradiation and by chemical reduction or oxidation. Moreover, al- though unbiased charge transfer in DNA nearly vanishes after 10 to 20 nm17,18, DNA still remains a promising candidate as an electronic component in molecular elec- tronics, e.g. as a short molecular wire19. Favoring geome- tries and base-pair sequences have still to be explored e.g. incorporation of sequences serving as molecular rectifiers, using non-natural bases or using the triplet acceptor an- thraquinone for hole injection20. Structural fluctuations could be another important factor which influences quan- tum transport through DNA molecular wires21. Finally, the carrier transfer rate through DNA can be manipu- lated by chemical modification22. On theoretical side, both ab initio calculations23 -- 29 and model Hamiltonians30 -- 38 try to interpret the diver- sity of experimental results and ascertain the underlying CT mechanism. The former can provide a more detailed description, but are currently limited to very short seg- ments, while the latter are much less detailed but allow- ing to address systems of realistic length, grasping hope- fully the underlying physics39. Here we study rather long B-DNA segments, hence we adopt the latter approach. Specifically, we employ two Tight-Binding (TB) ap- proaches. TB I is very simple, it is an approach at the base-pair (bp) level. We need the on-site energies of base pairs and the hopping integrals between successive base pairs. In other words, TB I is a wire model40. TB II is an approach at the single-base (sb) level. We need the on- site energies of bases and the hopping integrals between neighboring bases. We also include diagonal hoppings, in that sense, TB II is an extended ladder model39. The inclusion of diagonal hoppings is crucial in some cases as will become evident below. With these two TB models we study the electronic structure and hole or electron transfer in B-DNA monomer polymers and dimer poly- mers. This means that we call monomer a B-DNA base pair and study polymers made of N monomers, with rep- etition unit one or two monomers. To this end, we shall see below, we have to solve a system of M D ("matrix dimension") coupled equations for the time-independent problem, and a system of M D coupled 1st order differ- ential equations for the time-dependent problem. In TB I M D = N , while in TB II M D = 2N . In this article, we study HOMO and LUMO eigenspectra and the rele- vant Density of States (DOS) as well as the mean over time probabilities to find the carrier at each site, which is a base pair for TB I and a base for TB II. We are also interested in the frequency content of carrier move- ments, hence, we analyze the Fourier spectra, too. The pure mean transfer rate from a certain site to another describes the easiness of CT; it gives us a measure of how much of the carrier is transferred and also of how fast this process is. Our two TB approaches give coher- ent, complementary aspects of electronic properties and charge transfer in these B-DNA monomer polymers and dimer polymers. The rest of the paper is organized as follows: In Sec- tion II we delineate the basic theory behind the time- independent (Subsection II A) and the time-dependent (Subsection II B) problem. In Section III we discuss our results for eigenspectra and eigenvectors (Subsec- tion III A), the density of states (Subsection III B), the HOMO-LUMO gaps (Subsection III C), the mean over time probabilities to find the carrier at each site (Subsec- tion III D), the CT frequency content (Subsection III E), and the pure mean transfer rates (Subsection III F). In Section IV we state our conclusions. II. THEORY Let us begin with some notations. We call monomer a B-DNA base pair. We denote a system of two successive monomers by YX, according to the convention 2 for the B-DNA strands orientation. Xcompl (Ycompl) is the complementary base of X (Y). The base pair X- Xcompl is separated and twisted by 3.4 A and 36◦, re- spectively, relatively to the base pair Y-Ycompl, around the B-DNA growth axis. We call µ the monomer index, with µ = 1, 2, . . . , N , and σ the strand index (σ = 1 for the strand with 5′-3′ directionality, σ = 2 for the strand with 3′-5′ directionality). Further, we define the base in- dex β(µ, σ), β = 1, 2, . . . , 2N , according to the expression β = 2(µ − 1) + σ. Schematically, µ σ β 1 1 1 1 2 2 2 1 3 2 2 4 ... ... ... In this work, we study all possible periodic B-DNA seg- ments of the form YXYX. . . , consisting of N monomers, i.e., monomer polymers and dimer polymers. There are three types of such polymers: (type α′) poly(dG)-poly(dC), poly(dA)-poly(dT), (type β′) GCGC. . . , CGCG. . . , ATAT. . . , TATA. . . , and (type γ′) ACAC. . . , CACA. . . , TCTC. . . , CTCT. . . , AGAG. . . , GAGA. . . , TGTG. . . , GTGT. . . . We employ two Tight-Binding (TB) approaches to study the electronic structure and single carrier transfer in such B-DNA polymers, under the hypothesis that an extra hole or electron travels through HOMOs or LUMOs, re- spectively. (I) Within TB I (description at the base-pair level), we use the HOMO/LUMO on-site energies of base pairs and the HOMO/LUMO hopping integrals between successive base pairs. The TB parameters for TB I are the same as in Refs.17,18,41,42. (II) Within TB II (descrip- tion at the single-base level), we use the HOMO/LUMO on-site energies of bases and the HOMO/LUMO hop- ping integrals between (a) two successive bases on the same strand, (b) complementary bases that constitute a monomer, and (c) diagonally located bases of successive monomers, in the directions 5′-5′ and 3′ −3′. The TB pa- rameters for TB II are taken from Ref.43. In other words, within TB I, a monomer is considered as a single site, characterized by the index µ, while, within TB II, a base is considered as a single site, characterized by the index β. Below, we use a generic site index j, j = 1, 2, . . . , M D, where, j = µ and M D = N , for TB I, while, j = β and M D = 2N , for TB II. M D denotes "matrix dimension". σ=1 σ=2 3′ 5′ Y - Ycompl µ + 1 X - Xcompl ... µ ... 3′ 5′ A. Stationary States - Time-independent problem The HOMO/LUMO Hamiltonian of a given B-DNA segment can be written as (1) H = M D Xj=1 Es(j) jihj + X<j,j ′> ts(j,j ′)(cid:0) jihj′ + h.c.(cid:1), (2) where Es(j) is the HOMO/LUMO on-site energy of the j- th site [base pair (bp) for TB I or base (b) for TB II], and ts(j,j ′)(= ts(j ′,j)∗) is the HOMO/LUMO hopping integral between the sites j and j′. < j, j′ > denotes summa- tion over all relevant neighbors. The neighboring sites which are taken into account for each TB approach are described above. For TB I (wire model), the Hamiltonian can be written as HW = N Xµ=1 Ebp(µ) µihµ+(cid:18) N −1 Xµ=1 tbp(µ,µ+1) µihµ + 1+h.c.(cid:19). (3) For TB II (extended ladder model), the Hamiltonian can be written as HEL = Eb(β) βihβ (4) M D Xβ=1 +(cid:18) M D−2 Xβ=1 +(cid:18) M D−1 Xβ=1,odd +(cid:18) M D−3 Xβ=1,odd +(cid:18) M D−2 Xβ=2,even tb(β,β+2) βihβ + 2 + h.c.(cid:19) tb(β,β+1) βihβ + 1 + h.c.(cid:19) tb(β,β+3) βihβ + 3 + h.c.(cid:19) tb(β,β+1) βihβ + 1 + h.c.(cid:19), where the second term represents intra-strand, the third intra-base-pair, the fourth inter-strand 5′-5′ and the fifth inter-strand 3′-3′ hoppings. In the context of TB, we suppose that hjj′i = δjj ′ . The HOMO/LUMO state of the segment can be ex- pressed as DNAi = M D Xj=1 vj ji . (5) Substituting Eqs. (2) and (5) to the time-independent Schrodinger equation H DNAi = E DNAi , (6) we arrive to a system of M D coupled equations. Within TB I, the system is of the form Euµ = Ebp(µ)vµ + tbp(µ,µ+1)vµ+1 + tbp(µ,µ−1)vµ−1, (7) for µ even or odd, while, within TB II, the system is of the form Evβ =tb(β,β−2)vβ−2 + tb(β,β−1)vβ−1 + Eb(β)vβ+ tb(β,β+1)vβ+1 + tb(β,β+2)vβ+2 + tb(β,β+3)vβ+3, (8a) 3 for β odd, i.e., for the bases of strand 1, and Evβ =tb(β,β−3)vβ−3 + tb(β,β−2)vβ−2 + tb(β,β−1)vβ−1+ Eb(β)vβ + tb(β,β+1)vβ+1 + tb(β,β+2)vβ+2, (8b) for β even, i.e., for the bases of strand 2. Eqs. (7) and (8) are equivalent to the eigenvalue-eigenvector problem H~v = E~v, (9) where H is the hamiltonian matrix of order M D, com- posed of the TB parameters Es and ts, and ~v is the vector matrix composed of the coefficients vj. For the segments studied in this work, within TB I, H is ei- ther a tridiagonal symmetric Toeplitz matrix of order N for type α′ polymers, or a tridiagonal symmetric 2- Toeplitz matrix of order N for type β′ and type γ′ poly- mers. Within TB II, H is a heptadiagonal 4-Toeplitz matrix of order 2N , or, seen another way, a tridiagonal block Toeplitz matrix of order N 2 , with blocks of order 4. The diagonalization of H leads to the determination of the HOMO/LUMO eigenenergy spectra (eigenspec- tra), {Ek}, k = 1, 2, . . . , M D, for which we suppose that E1 < E2 < · · · < EM D, as well as to the determination of the occupation probabilities for each eigenstate, vjk2, where vjk is the j-th component of the k-th eigenvector. {vjk} are normalized, and their linear independence is checked in all cases. Having determined the eigenspectra, we can compute the Density of States (DOS), generally given by g(E) = M D Xk δ(E − Ek). (10) Changing the view of a B-DNA segment from one (e.g. top) to the other (e.g. bottom) side of the growth axis, reflects the hamiltonian matrix H of the segment on its main antidiagonal. This reflected Hamiltonian, H equiv, describes the equivalent polymer. H and H equiv are connected by the similarity transformation H equiv = P −1HP , where P (= P −1) is the unit antidiagonal ma- trix of order M D. Therefore, H and H equiv have iden- tical eigenspectra (hence the equivalent polymers' DOS is identical) and their eigenvectors are connected by vjk = vequiv (M D−j+1)k. For the types of B-DNA polymers studied in this work, equiv(YX. . . ) =(YcomplXcompl . . . , XcomplYcompl . . . , for odd N for even N (11) For example, for N odd, ACAC. . . ≡ TGTG. . . , while, for N even, ACAC. . . ≡ GTGT. . . . B. Time-dependent problem To describe the spatiotemporal evolution of an extra carrier (hole/electron), inserted or created (e.g. by oxi- dation/reduction) in a B-DNA segment, we consider the HOMO/LUMO state of the segment as DNA(t)i = M D Xj=1 Cj(t) ji , (12) where Cj(t)2 is the probability of finding the carrier at the j-th site at time t. Substituting Eqs. (2) and (12) to the time-dependent Schrodinger equation H DNA(t)i = i ∂ ∂t DNA(t)i , (13) we arrive at a system of M D coupled differential equa- tions of 1st order. With TB I, the system is of the form dCµ dt i = Ebp(µ)Cµ + tbp(µ,µ+1)Cµ+1 + tbp(µ,µ−1)Cµ−1, (14) for µ even or odd. With TB II, the system is of the form Suppose that the extra carrier is placed at the l-th site, that is Cl(0) = 1, Cj(0) = 0, ∀j 6= l. Then, 4 ~c = . (19) ul1 ... ulk ... ulM D     In other words, the coefficients ck are given by the row of the eigenvector matrix which corresponds to the site the carrier is initially placed at. In this work, within TB I, we choose l = 1, i.e., we initially place the extra carrier at the first monomer (initial condition), and, within TB II, we choose either l = 1 (initial condition 1 ) or l = 2 (initial condition 2 ), i.e., we initially place the extra carrier at each base of the first monomer. From Eq. (17) it follows that the probability to find the extra carrier at the j-th site of a B-DNA segment is i dCβ dt =tb(β,β−2)Cβ−2 + tb(β,β−1)Cβ−1 + Eb(β)Cβ+ tb(β,β+1)Cβ+1 + tb(β,β+2)Cβ+2 + tb(β,β+3)Cβ+3, (15a) Cj(t)2 = c2 kv2 jk +2 M D Xk=1 M D M D Xk=1 Xk′=1 k′<k ckck′ vjkvjk′ cos(2πfkk′ t), (20) for β odd, and where i dCβ dt =tb(β,β−3)Cβ−3 + tb(β,β−2)Cβ−2 + tb(β,β−1)Cβ−1+ Eb(β)Cβ + tb(β,β+1)Cβ+1 + tb(β,β+2)Cβ+2, (15b) for β even. Eqs. (14) and (15) are equivalent to a 1st order matrix differential equation of the form ~C(t) = − i  H ~C(t), (16) where ~C(t) is a vector matrix composed of the coefficients Cj(t), j = 1, 2, . . . , M D. Eq. (16) can be solved with the eigenvalue method, i.e., by looking for solutions of the form ~C(t) = ~ve− i  Et. Hence, Eq. (16) leads to the eigenvalue problem of Eq. (9), that is, H~v = E~v. Having determined the eigenvalues and eigenvectors of H, the general solution of Eq. (16) is ~C(t) = − i  E~ve− i  Et ⇒ ~C(t) = M D Xk=1 ck~vke− i  Ekt, (17) where the coefficients ck are determined from the initial conditions. In particular, if we define the M D × M D eigenvector matrix V , with elements vjk, then it can be shown that the vector matrix ~c, composed of the coeffi- cients ck, k = 1, 2, . . . , M D, is given by the expression ~c = V T ~C(0). (18) fkk′ = 1 Tkk′ = Ek − Ek′ h , ∀k > k′, (21) 2!(M −2)! = M(M −1) in CT is S =(cid:0)M are the frequencies (fkk′ ) or periods (Tkk′ ) involved in charge transfer. If M is the number of discrete eigenen- ergies, then, the number of different fkk′ or Tkk′ involved . If there are no degenerate eigenenergies (which holds for all cases stud- ied here, but e.g. does not hold in TB I for cyclic type α′ polymers18), then M = M D. If eigenenergies are sym- metric relative to some central value, then, S decreases (there exist degenerate fkk′ or Tkk′ ). Specifically, in that case, S = M 2 2(cid:1) = M! 4 , for even M and S = M 2−1 for odd M . From Eq. (20) it follows that the mean over time prob- 4 2 ability to find the extra carrier at the j-th site is DCj(t)2E = M D Xk=1 kv2 c2 jk. (22) Furthermore, from Eq. (20) it can be shown that the one-sided Fourier amplitude spectrum that corresponds to the probability Cj(t)2 is given by Fj(f ) = M D Xk=1 kv2 c2 jkδ(f ) + 2 M D M D Xk=1 Xk′=1 k′<k ckck′ vjkvjk′ δ(f − fkk′ ). (23) A quantity that can be defined to estimate the transfer rate, i.e., simultaneously, the magnitude of charge trans- fer and the time scale of the phenomenon, is the pure mean transfer rate kj ′j = DCj(t)2E tj ′j . (24) tj ′j is the mean transfer time, i.e., having placed the car- rier initially at site j′, the time it takes for the probabil- ity to find the extra carrier at site j, Cj(t)2, to become equal to its mean value, DCj(t)2E, for the first time. For the pure mean transfer rates it holds kj ′j = kjj ′ = kequiv (M D−j ′+1)(M D−j+1) = kequiv (M D−j+1)(M D−j ′ +1). (25) III. RESULTS A. Eigenspectra and occupation probabilities Let us start by saying that within TB II (TB I), we take the HOMO or LUMO eigenenergies of bases (base pairs) as the on-site energy of a hole or an electron on a base (base pair). Using the HOMO or LUMO ener- gies of the bases that constitute a base pair, we can es- timate the HOMO or LUMO energy of the base pair43. Specifically, supposing that ψbpi = C1 ψb1i + C2 ψb2i, and taking the time-independent Schrodinger equation H ψbpi = E ψbpi we find that the base pair eigenen- ergies are E1,2 = Eb1+Eb2 )2 + t2, where Eb1 and Eb2 are the on-site energies of the bases and ± q( Eb1−Eb2 2 2 1. type α′ polymers For TB I, an analytical expression for the eigenvalues of type α′ polymers exists18. All eigenvalues are symmetric around the on-site energy Ebp of the monomers and lie in the interval (cid:2)Ebp − 2(cid:12)(cid:12)tbp(cid:12)(cid:12), Ebp + 2(cid:12)(cid:12)tbp(cid:12)(cid:12)(cid:3). For odd N , the trivial eigenvalue, E = Ebp, exists. In the left column of Fig. 1 we present the calculated HOMO/LUMO eigen- spectra for an example of type α′ polymers, poly(dA)- poly(dT). For TB I, an analytical expression can also be found for the eigenvectors18. The eigenvectors (hence, the occupation probabilities, too) are eigenspectrum in- dependent 18, i.e., they do not depend on the TB param- eters Ebp, tbp. Furthermore, the occupation probabilities display palindromicity 18, i.e., the occupation probability of each eigenstate of the µ-th monomer is equal to that ). of the (N − µ + 1)-th monomer (uµk2 =(cid:12)(cid:12)u(N −µ+1)k(cid:12)(cid:12) For TB II, up to our knowledge, there are no ana- lytical expressions for eigenvalues and eigenvectors. As an example of type α′ polymers, we show, in the right column of Fig. 1, the calculated HOMO/LUMO eigen- 2 5 t = hψb1 Hψb2i is the intra-base-pair hopping integral i.e. between the two bases that constitute a base pair. However, due to the weak hydrogen bonding between the bases that constitute a base pair, t is very small43, of the order of 10 meV. As a result, practically, E1,2 ≈ Eb1, Eb2 (with accuracy of 1 meV). Hence, we make the Observa- tion: Approximately, the HOMO of the base pair is the highest HOMO of the two bases and the LUMO of the base pair is the lowest LUMO of the two bases. This is expressed in Table I, where we show all energies in eV with accuracy of 0.1 eV. Our numerical results for type TABLE I: On-site HOMO / LUMO energies of B-DNA bases and base pairs43. All energies are given in eV. −9.0 −4.9 4.1 −8.3 −4.4 3.9 base Adenine Thymine Guanine Cytosine Eb −8.8 H Eb −4.3 L 4.5 Eg base pair Ebp H Ebp L Eg −8.0 −4.5 3.5 −8.0 −4.5 −4.9 A-T −8.3 G-C 3.5 3.4 α′, β′, and γ′ polymers (cf. Figs. 1, 2, and 3), indicate that, as, increasing N , a polymer is formed, the energy eigenvalues are distributed around the on-site energies of the base pairs within TB I or the bases within TB II. Hence, the HOMO (LUMO) eigenspectrum of a given polymer within TB I corresponds to the upper (lower) part of its eigenspectrum within TB II. spectra of poly(dA)-poly(dT). The eigenvalues are dis- tributed in two subbands of different width, around the on-site energies of the bases. Furthermore, in accor- dance with the Observation, the upper (lower) subband of the HOMO (LUMO) eigenspectrum corresponds to the band calculated with TB I. For TB II, our nu- merical results for the eigenvectors indicate that, for β , while, for β even odd (strand 1), uβk2 ≈ (cid:12)(cid:12)u(2N −β)k(cid:12)(cid:12) uβk2 ≈ (cid:12)(cid:12)u(2N −β+2)k(cid:12)(cid:12) (strand 2), i.e., the occupa- tion probabilities of the eigenstates display approximate strand-palindromicity. For HOMO poly(dG)-poly(dC), a case where, according to the parameters used here43, the hopping integrals between diagonally located bases of successive monomers in the 3′-3′ and 5′-5′ directions are equal, strand palindomicity is strict. This also holds for all type α′ polymers, if our extended ladder model is reduced to a simple ladder model by neglecting 3′-3′ and 5′-5′ inter-strand interactions. 2 2 , LUMO poly(dA)-poly(dT) (TB I) -4.84 -4.86 -4.88 -4.90 -4.92 -4.94 -4.96 5 10 30 HOMO poly(dA)-poly(dT) (TB I) 20 25 15 -4.3 -4.4 -4.5 -4.6 -4.7 -4.8 -4.9 -5.0 -8.2 ) i V e ( s e g r e n e n e g E i -8.26 -8.28 -8.30 -8.32 -8.34 5 10 15 20 25 30 6 LUMO poly(dA)-poly(dT) (TB II) 5 10 30 HOMO poly(dA)-poly(dT) (TB II) 15 20 25 -8.4 -8.6 -8.8 -9.0 -9.2 N 5 10 15 20 25 30 FIG. 1: (Color online) An example of type α′ polymers: LUMO (first row) and HOMO (second row) eigenspectra of poly(dA)- poly(dT), for wire model (TB I, first column) and extended ladder model (TB II, second column). 2. type β′ polymers As far as equivalent polymers are concerned, for N even, reflection of the hamiltonian matrix H on its main antidiagonal leads to the same polymers, while for N odd, GCGC. . . ≡ CGCG. . . , ATAT. . . ≡ TATA. . . . For TB I, analytical expressions for the eigenvalues of type β′ polymers with N odd exist18. For N odd the eigenvalues can be expressed explicitly in terms of Chebyshev zeros44. All eigenvalues are symmetric around the on-site energy of the monomers, Ebp, and the trivial eigenvalue Ebp exists. The eigenvalues lie in the inter- val hEbp −pt2 1 + t2 2 + 2t1t2, Ebp +pt2 1 + t2 2 + 2t1t2i, where t1, t2 are the two different hopping integrals e.g., moving from the beginning to the end of the polymer, from-odd-to-even µ and from-even-to-odd µ, respectively. For N even, there is no explicit formula, although a recipe to produce the eigenvalues exists44. Our nu- merical results show that all eigenvalues are symmet- ric around the on-site energy Ebp of the monomers, and lie in the same interval as for N odd. The calculated HOMO/LUMO eigenspectrum for an example of type β′ polymers (GCGC. . . ), displaying all the above mentioned properties, is shown in the left column of Fig. 2. For TB I and N odd, analytical expressions for the eigenvectors exist45. These eigenvectors (hence, the oc- cupation probabilities, too) are partially eigenspectrum dependent 18, i.e., they depend on t1, t2 but not on Ebp. 2 Furthermore, for µ even, the occupation probability of each eigenstate of the µ-th monomer is equal to that of ), i.e., for N odd, the occupation probabilities of type β′ polymers display partial palindromicity 18. Finally, for N odd, equivalence leads to the property uµk2(YX. . .) = (XY. . .). For N even, we are aware of no an- alytical expressions for the eigenvectors, but our numeri- cal results show that the occupation probabilities display palindromicity 18, i.e., for each eigenstate, the occupation probability of the µ-th monomer is equal to that of the the (N − µ + 1)-th monomer (uµk2 = (cid:12)(cid:12)u(N −µ+1)k(cid:12)(cid:12) (cid:12)(cid:12)u(N −µ+1)k(cid:12)(cid:12) 2 ). (N − µ + 1)-th monomer (uµk2 =(cid:12)(cid:12)u(N −µ+1)k(cid:12)(cid:12) For TB II, up to our knowledge, there are no an- alytical expressions for eigenvalues and eigenvectors. As an example of type β′ polymers, we show in the right column of Fig. 2 the calculated HOMO/LUMO eigenspectra for GCGC. . . . The eigenvalues are dis- tributed in two subbands of different width, around the on-site energies of the bases. Moreover, in ac- cordance with the Observation, the upper (lower) sub- band of the HOMO (LUMO) eigenspectrum corresponds to the band calculated within TB I. For the TB II eigenvectors, for N odd, equivalence leads to the prop- (XY. . .), while for erty uβk2(YX. . .) = (cid:12)(cid:12)u(2N −β+1)k(cid:12)(cid:12) N even, uβk2 = (cid:12)(cid:12)u(2N −β+1)k(cid:12)(cid:12) In other words, the occupation probabilities of the eigenstates display base- palindromicity. 2 2 . 2 -4.48 -4.49 -4.50 -4.51 -4.52 LUMO GCGC... (TB I) 15 10 5 25 HOMO GCGC... (TB I) 20 ) i V e ( s e g r e n e n e g E i -7.94 -7.96 -7.98 -8.00 -8.02 -8.04 -8.06 5 10 15 20 25 30 7 LUMO GCGC... (TB II) 10 5 HOMO GCGC... (TB II) 15 20 25 30 5 10 15 20 25 30 -4.25 -4.30 -4.35 -4.40 -4.45 -4.50 -4.55 30 -7.8 -8.0 -8.2 -8.4 -8.6 -8.8 N FIG. 2: (Color online) An example of type β′ polymers: LUMO (first row) and HOMO (second row) eigenspectra of GCGC. . . , for wire model (TB I, first column) and extended ladder model (TB II, second column). 3. type γ′ polymers Here, ≡ TGTG. . . , CACA. . . the equivalent polymers are: for N odd, ACAC. . . ≡ GTGT. . . , CTCT. . . ≡ GAGA. . . , TCTC. . . ≡ AGAG. . . , and for N even, ACAC. . . ≡ GTGT. . . , CACA. . . ≡ TGTG. . . , CTCT. . . ≡ AGAG. . . , TCTC. . . ≡ GAGA. . . . e )/2. 1 + t2 1 + t2 o − Ebp o + Ebp Let us call Ebp e )/2 and ∆ = (Ebp ues lie in the interval [Σ −p∆2 + t2 p∆2 + t2 For TB I, analytical expressions for the eigenval- ues with N odd exist18. o(e) the on- site energy of monomers with µ odd (even), Σ = (Ebp Then, the eigenvalues include Ebp o , while the rest eigenval- 2 + 2t1t2, Σ + 2 + 2t1t2]. For N even, there is no explicit formula, although a recipe to produce the eigenvalues ex- ists44. Our numerical results show that all eigenvalues are symmetric around Σ, and lie in the same interval as for N odd. The calculated HOMO/LUMO eigenspec- trum for an example of type γ′ polymers (TCTC. . . ), displaying all the above mentioned properties, is shown in the left column of Fig. 3. For TB I and for N odd, analytical expressions can also be found for the eigenvectors46. The eigenvectors (hence, the occupa- tion probabilities, too) are eigenspectrum dependent 18, i.e., they depend on Ebp 2 , t1, t2. Furthermore, for µ even, the occupation probability of each eigenstate of 1 , Ebp 2 2 2 (XcomplYcompl . . . ). the µ-th monomer is equal to that of the (N − µ + 1)- ), i.e., for N , odd, the occupation probabilities of type γ′ polymers display partial palindromicity. For N odd, equivalence leads to (YcomplXcompl . . . ). For N even, up to our knowledge, no analytical expres- sions for the eigenvectors exist, but equivalence leads (XcomplYcompl . . . ). Our numerical results show that, for all µ, uµk2(YX. . .) = th monomer (uµk2 = (cid:12)(cid:12)u(N −µ+1)k(cid:12)(cid:12) uµk2(YX. . .) = (cid:12)(cid:12)u(N −µ+1)k(cid:12)(cid:12) to uµk2(YX. . .) =(cid:12)(cid:12)u(N −µ+1)k(cid:12)(cid:12) (cid:12)(cid:12)uµ(N −k+1)(cid:12)(cid:12) For TB II, there are no analytical expressions in the literature for eigenvalues and eigenvectors, as far as we know. The calculated HOMO/LUMO eigenspectra for an example of type β′ polymers (TCTC. . . ) are demonstrated in the right column of Fig. 3. The eigenvalues are distributed in four subbands of differ- ent width, around the on-site energies of the bases. Moreover, the two upper (lower) TB II subbands of the HOMO (LUMO) eigenspectrum correspond to the bands calculated with TB I. For the TB II eigenvectors, equivalence leads to the properties uβk2(YX. . .) = and (XcomplYcompl . . . ), for N even. Our numerical results indicate that there are no palindromic properties. (cid:12)(cid:12)u(2N −β+1)k(cid:12)(cid:12) uβk2(YX. . .) = (cid:12)(cid:12)u(2N −β+1)k(cid:12)(cid:12) in accordance with the Observation, (YcomplXcompl . . . ), for N odd, 2 2 2 LUMO TCTC... (TB I) LUMO TCTC... (TB II) 8 ) i V e ( s e g r e n e n e g E i -4.5 -4.6 -4.7 -4.8 -4.9 -7.9 -8.0 -8.1 -8.2 -8.3 -8.4 -4.3 -4.4 -4.5 -4.6 -4.7 -4.8 -4.9 30 5 -7.8 -8.0 -8.2 -8.4 -8.6 -8.8 -9.0 -9.2 N 5 15 10 25 HOMO TCTC... (TB I) 20 10 25 HOMO TCTC... (TB II) 15 20 30 5 10 15 20 25 30 5 10 15 20 25 30 FIG. 3: (Color online) An example of type γ′ polymers: LUMO (first row) and HOMO (second row) eigenspectra of TCTC. . . , for wire model (TB I, first column) and extended ladder model (TB II, second column). B. Density of States For TB I or TB II, the DOS can be determined directly by the eigenspectra (cf. Eq. (10)). It represents nicely the corresponding eigenspectral properties (cf. § III A). In Figs. 4, 5, and 6, we illustrate the numerically determined DOS for some representative examples of type α′, β′, and γ′ polymers, respectively, for N = 105, i.e., in the large N limit. We observe that, due to the fact that the eigenenergies become denser and denser as we approach the band or subband edges, van Hove Singularities (vHS) occur at the edges of each band or subband. We also notice that, in the large N limit, the polymer boundaries play insignificant role in the electronic structure, hence, for the same set of TB parameters, the polymers' DOS is essentially the same. For example, in the large N limit, either GCGC... or CGCG..., either N odd or N even have practically the same DOS. In some simpler cases, the DOS can be analytically obtained. For example, for type α′ polymers, within TB I, in the large N limit and using periodic boundaries, i.e., for cyclic type α′ polymers18, minigap, but for TB II there is a minigap of ≈ 0.545 eV; in accordance with the Observation, the upper subband of the HOMO band calculated with TB II, corresponds to the HOMO band calculated with TB I. The minigap is mainly due to the different HOMO on-site energies of the two bases (−8.0 eV for G, −8.8 eV for C)43. HOMO poly(dG)-poly(dC), N = 105, TB I -8.2 -8.15 -8.1 -8.05 -8 -7.95 -7.9 -7.85 -7.8 E (eV) HOMO poly(dG)-poly(dC), N = 105, TB II 60 50 40 30 ) 1 − V e ( ) E ( g N 20 10 0 20 15 10 ) 1 − V e ( ) E ( g N 5 g(E) = N πp4(tbp)2 − (E − Ebp)2 . (26) 0 -9 -8.8 -8.6 -8.4 E (eV) -8.2 -8 -7.8 For TB I, the numerically derived DOS for type α′ polymers (cf. Fig. 4) is in accordance with Eq. (26), because in the large N limit the boundary conditions play insignificant role. In Fig. 4, for TB I, there is no FIG. 4: (Color online) DOS for an example of type α′ poly- mers, poly(dG)-poly(dC) (N = 105, HOMO), for the base- pair (TB I, top) and the single-base (TB II, bottom) ap- proaches. A DOS example in type β′ polymers is shown in Fig. 5. For TB I, there is a small (≈ 0.004 eV) minigap. For TB II, there is a minigap of ≈ 0.200 eV; in accordance with the Observation, the lower subband of the LUMO band calculated with TB II corresponds to the LUMO band calculated with TB I. For TB II, there are two additional small (≈ 0.003 eV, 0.001 eV) minigaps, hardly noticeable at this scale. The underlined TB II minigap corresponds to the TB I minigap, also underlined. For TB II, apart from the vHSs at the subband edges, there is one vHS inside the second subband, hardly seen at this scale. LUMO CGCG..., N = 105, TB I HOMO CTCT...≡ AGAG..., N = 105, TB I 9 -8.3 -8.2 -8.1 -8 -7.9 E (eV) HOMO CTCT...≡ AGAG..., N = 105, TB II -9 -8.8 -8.6 -8.4 -8.2 -8 -7.8 E (eV) 100 ) 1 − V e ( ) E ( g N 80 60 40 20 0 -8.4 80 60 40 ) 1 − V e ( ) E ( g N 20 0 -9.2 600 500 400 300 ) 1 − V e ( ) E ( g N 200 100 0 -4.52 200 150 100 ) 1 − V e ( ) E ( g N 50 -4.515 -4.51 -4.505 -4.5 -4.495 -4.49 -4.485 -4.48 E (eV) LUMO CGCG..., N = 105, TB II FIG. 6: (Color online) DOS for an example of type γ′ poly- (N = 105, HOMO) for the mers, CTCT. . . ≡ AGAG. . . base-pair (TB I, top) and the single-base (TB II, bottom) approaches. possible monomer gaps. The decrease of the energy gap, as we move from monomer to polymer, is larger for type γ′ polymers. 0 -4.55 -4.5 -4.45 -4.4 E (eV) -4.35 -4.3 -4.25 FIG. 5: (Color online) DOS for an example of type β′ poly- mers, CGCG. . . (N = 105, LUMO), for the base-pair (TB I, top) and the single-base (TB II, bottom) approaches. A DOS example in type γ′ polymers is shown in Fig. 6. For TB I, there is a minigap a little greater than 0.340 eV, mainly due to the different HOMO on-site energies of the two base pairs (−8.0 eV for G-C, −8.3 eV for A-T)43. For TB II, four minibands are formed approximately around the HOMO on-site energies of the four bases (−9.0 eV for T, −8.8 eV for C, −8.3 eV for A and −8.0 eV for G)43, with three relevant minigaps (0.205 eV, 0.362 eV, 0.334 eV). Two of these minibands are very narrow. In accor- dance with the Observation, the higher two subbands of the HOMO band calculated with TB II correspond to the HOMO band calculated with TB I. The underlined TB II minigap corresponds to the TB I minigap, also underlined. 3.6 3.4 3.3 3.2 3.1 3.0 3.5 ) G-C monomer gap V e ( A-T monomer gap p a g O M U L - O M O H base-pair level single-base level 3.6 3.5 3.4 3.3 3.2 3.1 3.0 oly(d G )-p C ) oly(d p A )-p T) oly(d G C G C ..., C G ... C G oly(d p A C A A T A C ..., C A T A T..., T C A A ..., T C T C G ..., G T C T G T..., T T... G T C ..., G G A G ... A G A ..., A FIG. 7: (Color online) HOMO-LUMO gaps of type α′, β′, and γ′ polymers, for the base-pair (TB I, blue dots) and the single- base (TB II, purple dots) approaches. The red lines denote the HOMO-LUMO gaps of the two possible monomers. C. HOMO-LUMO gaps D. Mean over time probabilities In Fig. 7, we present the HOMO-LUMO energy gaps, in the large N limit, for all types of polymers. Both TB approaches predict similar gaps, in the range ≈ 3.04 - 3.42 eV. For TB I, the HOMO-LUMO gaps can also be derived analytically, from the maxima and minima of the HOMO and LUMO eigenspectra, respectively (cf. § III A). We also compare the polymer gaps with the two Within TB I, from Eq. (20) and the initial condition (carrier initially placed at the first monomer), it follows that the mean over time probability to find the extra carrier at the µ-th monomer is DCµ(t)2E = N Xk=1 v2 1kv2 µk. (27) Within TB II, from Eq. (20) and the initial condition 1 (carrier initially placed at the first base) or the initial condition 2 (carrier initially placed at the second base), it follows that the mean probability to find the extra carrier at the β-th base is 1.0 s e 0.8 i t i l i N = 12 DCβ(t)2E = v2 1kv2 βk v2 2kv2 βk 2N 2N Xk=1 Xk=1 . (28) 0.6 0.4 b a b o r P n a e M 0.2   10 = 12 = 11 = 10 = 9 = 8 = 7 = 6 = 5 = 4 = 3 = 2 = 1 1.0 0.8 0.6 0.4 0.2 0.0 From Eqs. (27) and (28), we conclude that the palin- dromicity and eigenspectrum (in)dependence properties for the occupation probabilities, presented in § III A, hold also for the mean over time probabilities. Fi- nally, for equivalent polymers it can be shown that in TB I hCN (t)2iYX. . . = hCN (t)2iequiv(YX. . . ), while in TB II hC2N (t)2iYX. . . = hC2N (t)2iequiv(YXYX. . . ) (for initial condition 1) and hC2N −1(t)2iYXYX. . . = hC2N −1(t)2iequiv(YXYX. . . ) (for initial condition 2). C ) H 0.0 oly(d G )-p oly(d p oly(d p C ) L oly(d G )-p A )-p oly(d p T) H oly(d T) L oly(d A )-p oly(d p FIG. 8: (Color online) Type α′ polymers. TB I and initial condition (extra carrier initially at the first base pair). Mean over time probabilities to find an extra hole (HOMO) or elec- tron (LUMO) at each base pair. Here N = 12. s e i t i l i b a b o r P n a e M 1.0 0.8 0.6 0.4 0.2 0.0 oly(d G )-p oly(d p N = 12 N = 12 = 1 = 2 = 12 = 11 = 10 = 9 = 8 = 7 = 6 = 5 = 4 = 3 = 2 = 1 C ) L oly(d G )-p C ) H oly(d p A )-p oly(d p T) H oly(d A )-p oly(d p T) L oly(d oly(d G )-p C ) H oly(d p oly(d G )-p T) H oly(d A )-p C ) L oly(d p A )-p oly(d p T) L oly(d oly(d p 1.0 0.8 0.6 0.4 0.2 0.0 FIG. 9: (Color online) Type α′ polymers. TB II and either initial condition 1 (left) or initial condition 2 (right), i.e., extra carrier initially at the first or the second base of the first base pair, respectively. Mean over time probabilities to find an extra hole (HOMO) or electron (LUMO) at each base. Here N = 12. 1. type α′ polymers In Figs. 8-9, we show an example (for N = 12) of our numerical results for the mean over time probabilities to find an extra hole or electron at (I) each base pair according to TB I and the initial condition (Fig. 8), and (II) each base according to TB II and the initial condition 1 or the initial condition 2 (Fig. 9), for type α′ polymers. For TB I, the mean over time probabilities to find the carrier at a specific monomer display palindromicity and eigenspectrum independence 18. Specifically, it can ana- lytically be shown that 3 DC1(t)2E =DCN (t)2E = DC2(t)2E = · · · =DCN −1(t)2E = 2(N + 1) , ∀N ≥ 2, (29a) 1 N + 1 , ∀N ≥ 3. (29b) For TB II, the mean over time probabilities to find the carrier at a specific base display approximate strand- palindromicity. Moreover, adding the mean probabili- ties of the bases that constitute each monomer, it fol- lows that the mean probabilities to find the carrier at a specific monomer are approximately palindromes and almost identically equal, for all cases, to the mean prob- abilities within TB I, which are strictly palindromes (cf. Eq. (29)). This quantitative agreement suggests that the eigenspectrum independence predicted within the sim- pler TB I approach, leads to essentially the same re- sults as those derived by the more complicated TB II approach. In Fig. 9 we observe that, within TB II, the carrier moves almost exclusively through the strand it was initially placed at, i.e. carrier movement is mainly of intra-strand character. Furthermore, within TB II, our results for the two initial conditions are in complete agreement. 2. type β′ polymers In Figs. 10-11 we present examples of our numerical results for type β′ polymers, for the mean over time prob- abilities to find an extra hole or electron (I) at each base pair according to TB I and the initial condition (Fig. 10), and (II) at each base according to TB II and the initial condition 1 or the initial condition 2 (Fig. 11). For TB I, the mean probabilities to find the carrier at a specific monomer display18 partial eigenspectrum depen- dence (i.e., dependence on the hopping parameters but not on the on-site energy), partial palindromicity (i.e., only for even µ) for N odd and palindromicity (i.e., for all µ) for N even. For TB II, for N even, the mean probabilities to find the carrier at a specific base display base-palindromicity. Moreover, adding the mean probabilities of the bases that constitute each base pair, the mean probabilities to find the carrier at a specific base pair are palindromes, in accordance with the prediction of TB I. In Fig. 11, we observe that within TB II, the carrier moves prefer- E. Charge transfer frequency content 4 for N odd and N 2 For TB I, for type α′ and β′ polymers, all eigenvalues are symmetric around the on-site energy of the base pairs. Hence, the total number of frequencies involved in charge transfer is N 2−1 4 for N even. For type γ′ polymers with N even, the eigenvalues are symmetric around Ebp , hence the total number of frequencies is N 2 4 , too. For type γ′ polymers with N odd, the eigen- values include Ebp 1 and the total number of frequencies is (N −1)(N +3) . For TB II, there are no symmetries like those mentioned for TB I, hence the total number of fre- 1 +Ebp 2 4 2 11 ably through the bases that are identical with the one it was initially placed at, in other words it moves crosswise through identical bases, i.e., carrier movement is mainly of inter-strand character. If we add or subtract a monomer, i.e. For N odd, both TB approaches show that there are some cases, in which the carrier hardly moves from its initial site. for N even, both TB approaches show that a large percent- age of the carrier is transferred at the end monomer. Furthermore, both TB approaches show that the mean probability to find the carrier at the last monomer is generally bigger for N even than for N odd. 3. type γ′ polymers In Figs. 12 and 13, we present examples of our numer- ical results for the mean over time probabilities to find an extra hole or electron (I) at each base pair according to TB I and the initial condition (Fig. 12), and (II) at each base according to TB II and the initial condition 1 or the initial condition 2 (Fig. 13), for type γ′ polymers. = (27) leads to identical mean probabilities for (i) TCTC. . . and AGAG. . . , (iii) and GAGA. . . , (ii) CTCT. . . ACAC. . . and TGTG. . . . that (XcomplYcompl . . . ), uµk2(YX. . .) for all µ, Eq. (cid:12)(cid:12)uµ(N −k+1)(cid:12)(cid:12) and GTGT. . . , and (iv ) CACA. . . given TB In I, 2 In Fig. 12 we observe that within TB I, the carrier moves preferably through the monomers that are iden- tical with the one it was initially placed at, i.e., from the first monomer to the third, and so forth. Within TB II, the carrier moves preferably through the bases that are identical with the one it was initially placed at, i.e., it moves through the same strand from the one or the other base of the first monomer to the identical base of the third monomer, and so forth, i.e., carrier movement is mainly of inter-strand character. Both TB approaches show that the mean probability to find the carrier at the last monomer is bigger for N odd than for N even, cf. Figs. 12-13. quencies for all types of polymers (α′, β′, γ′) is N (2N −1). From Eq. (23) it follows that all the palindromicity and equivalence properties presented in § III D for the mean over time probabilities, < Cj(t)2 >, hold for the Fourier spectra, Fj(f ), too. In the following Subsections, we focus on the Fourier spectra that correspond to charge transfer from the first to the last monomer i.e. on F1(f ) and FN (f ) for TB I, and on F1(f ), F2(f ), F2N −1(f ) and F2N (f ) for TB II. Both TB approaches show that the frequency content is mainly in the THz domain, cf. Figs. 14, 15, 16, 17, 18, 19. N = 12 N = 13 1.0 s e 0.8 i t i l i b a b o r P n a e M 0.6 0.4 0.2 0.0 G C G C ... H C ... L C G C G C G G ... H C G C G ... L T... H A T A T... L A T A T A T A... H A... L T A T 1.0 0.8 0.6 0.4 0.2 0.0 = 12 = 11 = 10 = 9 = 8 = 7 = 6 = 5 = 4 = 3 = 2 = 1 1.0 s e 0.8 i t i l i b a b o r P n a e M 0.6 0.4 0.2 0.0 G C G C ... H C ... L C G C G C G G ... H C G C G ... L T... H A T A T... L A T A T A T A... H A... L T A T 12 1.0 0.8 0.6 0.4 0.2 0.0 = 13 = 12 = 11 = 10 = 9 = 8 = 7 = 6 = 5 = 4 = 3 = 2 = 1 FIG. 10: (Color online) Type β′ polymers. TB I and initial condition (extra carrier initially at the first base pair). Mean over time probabilities to find an extra hole (HOMO) or electron (LUMO) at each base pair. N even (here N = 12, left) or N odd (here, N = 13, right). s e i t i l i b a b o r P n a e M s e i t i l i b a b o r P n a e M 1.0 0.8 0.6 0.4 0.2 0.0 G C G 1.0 0.8 0.6 0.4 0.2 0.0 G C G C ... H C ... H G G C C G G C ... L C ... L G ... H G ... H C C G G C C C C G G C C N = 12 N = 12 = 2 = 1 = 2 = 1 = 12 = 11 = 10 = 9 = 8 = 7 = 6 = 5 = 4 = 3 = 2 = 1 C G = 13 = 12 = 11 = 10 = 9 = 8 = 7 = 6 = 5 = 4 = 3 = 2 = 1 C G G G C ... H C ... H G G C C G G C ... L C ... L G ... H G ... H C C G G C C C C G G C C G ... L T... H T... L A T A T A T A... H T A T A... L A T A N = 13 G ... L T... H T... L A T A T A T A... H T A T A... L A T A G ... L T... H T... L A T A T A T A... H T A T A... L A T A N = 13 G ... L T... H T... L A T A T A T A... H T A T A... L A T A 1.0 0.8 0.6 0.4 0.2 0.0 1.0 0.8 0.6 0.4 0.2 0.0 FIG. 11: (Color online) Type β′ polymers. TB II and either initial condition 1 (left) or initial condition 2 (right), i.e., extra carrier initially at the first or the second base of the first base pair, respectively. Mean over time probabilities to find an extra hole (HOMO) or electron (LUMO) at each base. N even (upper panels, here N = 12) or N odd (lower panels, here N = 13). N = 12 1.0 s e 0.8 i t i l i 0.6 b a b o r P n a e M 0.4 0.2 0.0 C ..., G T C T G C A T A... H C ..., G T A... L T..., A G T A C C G ... H T..., A C G ... L C ..., G A T... H C ..., G A T... L A..., T C G A T C G C T A T A G C A C G A A T G C G ... H A..., T C G ... L T G 1.0 0.8 0.6 0.4 0.2 0.0 = 12 = 11 = 10 = 9 = 8 = 7 = 6 = 5 = 4 = 3 = 2 = 1 1.0 N = 13 0.8 s e i t i l i 0.6 b a b o r P n a e M 0.4 0.2 0.0 C ..., G T C T G C A T A ... H C ..., G T A ... L T..., A C G ... H T..., A C A G T C A C G A G ... L C ..., G A G T A C T... H C ..., G A T... L A ..., T C G ... H A ..., T C T G A C G A T C G ... L T G T A G C 13 1.0 0.8 0.6 0.4 0.2 0.0 = 13 = 12 = 11 = 10 = 9 = 8 = 7 = 6 = 5 = 4 = 3 = 2 = 1 FIG. 12: (Color online) Type γ′ polymers. TB I and initial condition (extra carrier initially at the first base pair). Mean over time probabilities to find an extra hole (HOMO) or electron (LUMO) at each base pair. N even (here N = 12, left) or N odd (here N = 13, right). 1.0 0.8 0.6 0.4 0.2 s e i t i l i b a b o r P n a e M 0.0 C ... H T C T T C 1.0 0.8 0.6 0.4 0.2 0.0 C ... H T C T T C T s e i t i l i b a b o r P n a e M T N = 12 N = 12 = 1 = 2 = 12 = 11 = 10 = 9 = 8 = 7 = 6 = 5 = 4 = 3 = 2 = 1 A... L G A A... H G A G A... L A G A G ... H A G A G ... L G T G T... H G T G T... L T G T G ... H T G T G ... L C ... H T C T C ... L C T C T... H C T C T... L A C A C ... H A C A C ... L C A C A... H C A C T C T N = 13 = 1 = 2 = 13 = 12 = 11 = 10 = 9 = 8 = 7 = 6 = 5 = 4 = 3 = 2 = 1 A... L G A A... H G A G A... L A G A G ... H A G A G ... L G T G T... H G T G T... L T G T G ... H T G T G ... L C ... H T C T C ... L C T C T... H C T C T... L A C A C ... H A C A C ... L C A C A... H C A C T C T C ... L C T C T... H C T C T... L A C A C ... H A C A C ... L C A C A... H C A C C ... L C T C T... H C T C T... L A C A C ... H A C A C ... L C A C A... H C A C G G G G A... L G A A... H G A G A... L A G A G ... H A G A G ... L G T G T... H G T G T... L T G T G ... H T G T G ... L N = 13 A... L G A A... H G A G A... L A G A G ... H A G A G ... L G T G T... H G T G T... L T G T G ... H T G T G ... L 1.0 0.8 0.6 0.4 0.2 0.0 1.0 0.8 0.6 0.4 0.2 0.0 FIG. 13: (Color online) Type γ′ polymers. TB II and either initial condition 1 (left) or initial condition 2 (right) , i.e., extra carrier initially at the first or the second base of the first base pair, respectively. Mean over time probabilities to find an extra hole (HOMO) or electron (LUMO) at each base. N even (upper panels, here N = 12) or N odd (lower panels, here N = 13). 1. type α′ polymers efficient than initial condition 2 for hole transfer, cf. also Figs. 12 and 13. 14 Within TB I the main frequencies are in the range ≈ 0.3 - 97 THz. Within TB II, they are in the range ≈ 0.1 - 110 THz. The main frequency content is between far-infrared (FIR) and mid-infrared (MIR). As an exam- ple, we show in Fig. 14 (TB I and initial condition) and Fig. 17 (TB II and initial condition 1 or initial condition 2 ), the Fourier spectra, at the first and the last monomer, of an extra hole in poly(dA)-poly(dT) with N = 20. In Fig. 14 we observe that the Fourier amplitudes for the first and the last monomer are approximately equal, mir- roring the efficient hole transfer in poly(dA)-poly(dT), cf. also Fig. 8. Inspection of Fig. 17 leads to the same conclusion. Additionally, Fig. 17 underlines the intra- strand character of carrier transfer and shows that initial conditions 1 and 2 lead to similar form of Fourier spectra. 2. type β′ polymers Within TB I, the main frequencies are in the range ≈ 0.01 - 40 THz, i.e., between microwaves (MW) and MIR. Within TB II, they are in the range ≈ 0.01 - 210 THz, i.e., between the MW and near-infrared NIR. As an exam- ple, we show in Fig. 15 (TB I and initial condition) and Fig. 18 (TB II and initial condition 1 or initial condition 2 ), the Fourier spectra, at the first and the last monomer, of an extra electron in ATAT. . . with N = 14. In Fig. 15 we observe that the Fourier amplitudes for the first and the last monomer are approximately equal, mirroring the finally large electron transfer in ATAT. . . for N even, cf. also Fig. 10. However, this large transfer is very slow, its main frequency is very small but with a large amplitude. The same conclusions can be drawn from Fig. 18, where we can additionally observe the inter-strand character of charge transfer and that initial conditions 1 and 2 lead to similar form of Fourier spectra. 3. type γ′ polymers Within TB I, the main frequencies are in the range ≈ 0.4 GHz - 40 THz, i.e., between radiowaves (RW) and MIR. Within TB II, they are in the range ≈ 0.02 - 190 THz, i.e., between MW and FIR. As an example, we show in Fig. 16 (TB I and initial condition) and Fig. 19 (TB II and initial condition 1 or initial condition 2 ), the Fourier spectra, at the first and the last monomer, of an extra hole in TCTC. . . with N = 21. In Fig. 16 we ob- serve that the Fourier amplitudes for the first monomer are much larger than the ones for the last monomer, mir- roring the inefficient hole transfer in TCTC. . . for N odd, cf. also Fig. 12. In Fig. 19, we can additionally observe the intra-strand character of charge transfer and that initial conditions 1 and 2 lead to somehow different form of Fourier spectra, initial condition 1 being more HOMO poly(dA)-poly(dT), N = 20 FA for C1(t)2 FA for CN(t)2 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 e d u t i l p m A r e i r u o F 10-8 10-7 10-6 10-5 10-4 10-3 10-2 10-1 100 101 102 103 f (THz) FIG. 14: (Color online) Type α′ polymers, here poly(dA)- poly(dT), N = 20. TB I and initial condition (extra carrier initially at the first base pair). Hole transfer Fourier spectra at the first and the last monomer. LUMO ATAT..., N = 14 FA for C1(t)2 FA for CN(t)2 HOMO TCTC..., GAGA..., N = 21 FA for C1(t)2 FA for CN(t)2 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 e d u t i l p m A r e i r u o F 10-8 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 e d u t i l p m A r e i r u o F 10-8 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 10-8 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 10-8 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 10-8 10-7 10-6 10-5 10-4 10-3 10-2 10-1 100 101 102 103 f (THz) FIG. 15: (Color online) Type β′ polymers, here ATAT. . . , N = 14. TB I and initial condition (extra carrier initially at the first base pair). Electron transfer Fourier spectra at the first and the last monomer. 10-7 10-6 10-5 10-4 10-3 10-2 10-1 100 101 102 103 f (THz) FIG. 16: (Color online) Type γ′ polymers, here TCTC. . . , N = 21. TB I and initial condition (extra carrier initially at the first base pair). Hole transfer Fourier spectra at the first and the last monomer. HOMO poly(dA)-poly(dT), N = 20 15 initial condition 1 FA for C1(t)2 FA for C2(t)2 FA for C2N-1(t)2 FA for C2N(t)2 initial condition 2 FA for C1(t)2 FA for C2(t)2 FA for C2N-1(t)2 FA for C2N(t)2 10-7 10-6 10-5 10-4 10-3 10-2 10-1 100 101 102 103 10-7 10-6 10-5 10-4 10-3 10-2 10-1 100 101 102 103 FIG. 17: (Color online) Type α′ polymers, here poly(dA)-poly(dT), N = 20. TB II and either initial condition 1 (left) or initial condition 2 (right). Hole transfer Fourier spectra at the first and the last monomer. 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 e d u t i l p m A r e i r u o F 10-8 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 e d u t i l p m A r e i r u o F 10-8 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 e d u t i l p m A r e i r u o F 10-8 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 10-8 f (THz) 10-8 f (THz) 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 10-8 f (THz) FIG. 18: (Color online) Type β′ polymers, here ATAT. . . , N = 14. TB II and either initial condition 1 (left) or initial condition 2 (right). Electron transfer Fourier spectra at the first and the last monomer. HOMO TCTC..., N = 21 initial condition 1 FA for C1(t)2 FA for C2(t)2 FA for C2N-1(t)2 FA for C2N(t)2 LUMO ATAT..., N = 14 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 initial condition 2 FA for C1(t)2 FA for C2(t)2 FA for C2N-1(t)2 FA for C2N(t)2 10-7 10-6 10-5 10-4 10-3 10-2 10-1 100 101 102 103 10-7 10-6 10-5 10-4 10-3 10-2 10-1 100 101 102 103 initial condition 1 FA for C1(t)2 FA for C2(t)2 FA for C2N-1(t)2 FA for C2N(t)2 initial condition 2 FA for C1(t)2 FA for C2(t)2 FA for C2N-1(t)2 FA for C2N(t)2 10-7 10-6 10-5 10-4 10-3 10-2 10-1 100 101 102 103 10-7 10-6 10-5 10-4 10-3 10-2 10-1 100 101 102 103 FIG. 19: (Color online) Type γ′ polymers, here TCTC. . . , N = 21. TB II and either initial condition 1 (left) or initial condition 2 (right). Hole transfer Fourier spectra at the first and the last monomer. F. Pure mean transfer rates either within TB I or within TB II. In the following subsections we focus on pure mean transfer rates between the first and the last monomer, 16 1. type α′ polymers 3. type γ′ polymers As a characteristic example, we present in Fig. 20 the hole pure mean transfer rates for poly(dG)-poly(dC), from the first to the last monomer, either within TB I or within TB II. Specifically, (I) for TB I we illustrate k1,N on the left panel, and (II) for TB II we illustrate k1,2 = k2,1, k1,2N −1 = k2,2N and k1,2N = k2,2N −1 on the right panel. We have already noticed in § III D 1 that, within TB II, carrier transfer is almost exclu- sively of intra-strand character. Hence, within TB II, k1,2N −1 = k2,2N are the largest transfer rates. Com- paring k1,N for TB I with k1,2N −1 = k2,2N for TB II, we observe an excellent agreement, both qualitatively and quantitatively. Within TB II, the intra-base-pair rates k1,2 = k2,1 are small and the inter-strand rates k1,2N = k2,2N −1 insignificant. Increasing N , the intra- strand transfer rates k1,2N −1 = k2,2N decrease reach- ing gradually the level of the the intra-base-pair rates k1,2 = k2,1, at which point, finally, charge transfer along the polymer is insignificant. Increasing N , the insignif- icant inter-strand rates k1,2N = k2,2N −1 also gradually decrease further. 2. type β′ polymers We have already mentioned (cf. § III D 2 and III E 2) that both TB approaches predict that for some cases of type β′ polymers, for N even, the carrier is transferred at a large percentage to the last monomer but the transfer is very slow. Such a case is presented in Fig. 21. Specif- ically, we show the electron pure mean transfer rates for ATAT. . . , from the first to the last monomer, either within TB I or within TB II. Specifically, (I) for TB I, we illustrate k1,N (left), and (II) for TB II, we illustrate the largest transfer rates (right). We have already demon- strated in § III D 2 that, within TB II, the extra carrier is transferred almost exclusively crosswise, through iden- tical bases. Hence, for TB II, the largest transfer rates are k1,2N −1 and k2,2N for N odd, and k1,2N and k2,2N −1 for N even. We depict these largest transfer rates in Fig. 21 (right). In other cases of type β′ polymers the pure mean transfer rates fall over N in a different man- ner, somehow similar to the behavior of type γ′ polymers, which is shown in § III F 3. As a characteristic example, we present in Fig. 22 the hole pure mean transfer rates for TCTC. . . , from the first to the last monomer, either within TB I or within TB II. Specifically, (I) for TB I we illustrate k1,N on the left panel, and (II) for TB II we illustrate k1,2N −1 and k2,2N on the right panel. We have already men- tioned in § III D 3 that, within TB II, the extra carrier is transferred almost exclusively through the strand it was initially placed at, i.e., for type γ′ polymers the charge transfer is mainly of intra-strand character. Hence, for TB II, we show k1,2N −1 and k2,2N which are the largest transfer rates. We have demonstrated §III F 1 that, for type α′ polymers, k1,2N −1 = k2,2N . As shown in Fig. 22, this does not hold for type γ′ polymers. 4. Pure mean transfer rate fits Finally, we compare the results of our two TB ap- proaches by performing the exponential fits k = k0e−βd and k = A + k0e−βd, where d = (N − 1) × 3.4 A is the charge transfer distance, as well as the power-law fit, 0N −η. Our results for TB I have already been pre- k = k′ sented in Figs. 8-9 of Ref.18. For TB II, we again focus on the pure mean transfer rates between the bases of the initial and the final monomer for which carrier transfer is significant. The conclusions are similar to those within TB I. The fits are considerably improved if polymers with N odd and N even are fitted separately. Moreover, the power-law fits are generally better, in terms of correla- tion coefficients. Our results for the exponent η of the power-law fits, within TB II, are presented in Fig. 23. Our results confirm the statement that the fall of k as a function of N becomes generally steeper as the intri- cacy of the energy structure increases, i.e., from type α′ to type β′ and further to type γ′ polymers18. This con- clusion also holds for the exponential fits, which are not presented here. Furthermore, both TB I and TB II show that there is perfect agreement between our results for β and η for all type α′ polymers. This leads to the conclu- sion that although the interaction strength (as reflected in the hopping integrals) is different in each case of type α′ polymers leading to different values of k, the way k falls over N or d is the same. 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 10-8 ) z H P ( k HOMO poly(dG)-poly(dC) (TB I) k1,N 0 5 10 15 N 20 25 30 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 10-8 ) z H P ( k 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 10-8 HOMO poly(dG)-poly(dC) (TB II) k1,2 = k2,1 k1,2N-1 = k2,2N k1,2N = k2,2N-1 0 5 10 15 N 20 25 30 17 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 10-8 FIG. 20: (Color online) Type α′ polymers, here poly(dG)-poly(dC). Hole pure mean transfer rates (I) k1,N for TB I (left), and (II) k1,2 = k2,1, k1,2N −1 = k2,2N and k1,2N = k2,2N −1 for TB II (right). 100 10-2 10-4 10-6 10-8 10-10 10-12 ) z H P ( k LUMO ATAT... (TB I) k1,N 0 5 10 15 N 20 25 30 100 10-2 10-4 10-6 10-8 10-10 10-12 100 10-2 10-4 10-6 10-8 10-10 10-12 ) z H P ( k LUMO ATAT... (TB II) k1,2N-1 k1,2N k2,2N-1 k2,2N 0 5 10 15 N 20 25 30 100 10-2 10-4 10-6 10-8 10-10 10-12 FIG. 21: (Color online) Type β′ polymers, here is a case where, for N even, the carrier is transferred at a large percentage to the last monomer but the transfer is very slow: Electron pure mean transfer rates in ATAT. . . , either k1,N within TB I (left) or the largest transfer rates in TB II, i.e., k1,2N −1 for N odd, k1,2N for N even, k2,2N for N odd, k2,2N −1 for N even. 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 10-8 ) z H P ( k HOMO TCTC... (TB I) k1,N 0 5 10 15 N 20 25 30 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 10-8 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 10-8 ) z H P ( k HOMO TCTC... (TB II) k1,2N-1 k2,2N 0 5 10 15 N 20 25 30 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 10-8 FIG. 22: (Color online) Type γ′ polymers, here TCTC. . . . Hole pure mean transfer rates k1,N for TB I (left), and k1,2N −1, k2,2N for TB II (right). IV. CONCLUSION We employed two Tight-Binding approaches to ex- amine time-independent and time-dependent aspects of the electronic structure and carrier transfer in B-DNA monomer polymers (type α′) and dimer polymers (type β′ and type γ′). We used a simplistic wire model (TB I) where a carrier is initially located at a base pair (called also a monomer in this article) and then moves to the next or to the previous base pair, as well as a more de- tailed extended ladder model (TB II) where the carrier is initially located at a base and then moves to all possible neighboring bases including diagonally located ones. The inclusion of diagonal hoppings is crucial for type β′ poly- mers where carrier transfer is mainly of inter-strand char- acter. The time-dependent and the time-independent problems involve diagonalization of matrices with ma- trix dimension M D = N for TB I and M D = 2N for TB II. The two TB approaches give coherent, complemen- tary aspects of electronic properties and charge transfer in B-DNA monomer polymers and dimer polymers. For the time-independent problem, we studied the HOMO and the LUMO eigenspectra and the occupa- tion probabilities, the Density of States and the HOMO- LUMO gap. The upper (lower) subband of the HOMO (LUMO) eigenspectrum calculated with TB II corre- sponds to the band calculated with TB I. The occupation probabilities within TB I and TB II show various degrees of palindromicity and eigenspectrum (in)dependence of the probabilities to find the carrier at a site. The DOS displays nice van Hove singularities at the (sub)band edges, while the numerically calculated DOS for simple cases agrees with the analytical solution. As expected, 18 the polymer HOMO-LUMO gaps are smaller than the HOMO-LUMO gaps of the two possible monomers, reaching a level of 3.4 to 3.0 eV. The smallest HOMO- LUMO gaps occur for type γ′ polymers. For the time-dependent problem we investigated the mean over time probabilities to find the carrier at each site (base pair for TB I and base for TB II), the Fourier spectra and the pure mean transfer rates from a certain site to another. The mean over time probabilities illus- trate clearly the basically intra-strand character of car- rier transfer in type α′ and type γ′ polymers. However, while in type α′ polymers the carrier moves successively through all bases of the same strand, in type γ′ poly- mers the carrier moves through the bases that are iden- tical with the one it was initially placed at, i.e., it moves through the same strand from the one or the other base of the first monomer to the identical base of the third monomer, and so forth. Carrier transfer is basically of inter-strand character in type β′ polymers. The Fourier spectra give us a nice representation of the frequency con- tent of charge transfer. Both TB approaches show that this frequency content is mainly in the THz domain, the details depend on the type of polymers and the TB ap- proach used. The pure mean transfer rates k show both how fast carrier transfer is and how much of the carrier is transferred from the initial site to the final site. The k(N ) fits are considerably improved if polymers with N odd and N even are fitted separately. Additionally, the power-law fits are generally better, in terms of correlation coefficients. Our results confirm the statement that the fall of k as a function of N becomes generally steeper as the intricacy of the energy structure increases, i.e., from type α′ to type β′ and further to type γ′ polymers. URL: http://users.uoa.gr/~csimseri/physics_of_nanostructures_and_biomaterials.html ∗ Current affiliation: University of Copenhagen, Niels Bohr Institute, Blegdamsvej 17, DK-2100 Copenhagen, Den- mark. † Current affiliation: Georg-August-Universitat Gottingen, Fakultat fur Physik, Friedrich-Hund-Platz 1 D-37077 Gottingen, Germany. ‡ Electronic address: [email protected]; 1 C.C. Page, C.C Moser and P. Leslie Dutton, Mechanism for electron transfer within and between proteins, Current Opinion in Chemical Biology 7, 551 (2003). 2 H.B. Gray and J.R. Winkler, Electron flow through met- alloproteins, Biochimica et Biophysica Acta 1797, 1563 (2010). 3 C.C. Moser, J.L. Ross Anderson, P. Leslie Dutton, Guide- lines for tunneling in enzymes, Biochimica et Biophysica Acta 1797, 1573 (2010). 4 J. M. Art´es, M. L´opez-Mart´ınez, I. D´ıez-P´erez, F. Sanz and P. Gorostiza, Nanoscale charge transfer in redox proteins and DNA: Towards biomolecular electronics, Electrochim- ica Acta 140, 83 (2014). 5 A.M. Kannan, V. Renugopalakrishnan, S. Filipek, P. Li, G.F. Audette, and L. Munukutla, Bio-Batteries and Bio- Fuel Cells: Leveraging on Electronic Charge Transfer Pro- teins, J. Nanosci. Nanotechnol. 9, 1665 (2009). 6 P.J. Dandliker, R. Erik Holmlin, J.K. Barton, Oxidative Thymine Dimer Repair in the DNA Helix, Science 275, 1465 (1997). 7 S.R. Rajski, B.A. Jackson, J.K. Barton, DNA repair: mod- els for damage and mismatch recognition, Mutation Re- search 447, 49 (2000). 8 B. Giese, Electron transfer through DNA and peptides, Bioorganic & Medicinal Chemistry 14, 6139 (2006). 9 C.-T. Shih, Y.-Y. Cheng, S.A. Wells, C.-L. Hsu, R.A. Romer, Charge transport in cancer-related genes and early carcinogenesis, Computer Physics Communications 182, 36 (2011). 10 H.-W. Fink and C. Schonenberger, Electrical conduction 19 k1,2N-1 k2,2N k1,2N k2,2N-1 k1,2N-1 k2,2N k1,2N-1 k2,2N 14 12 10 8 6 4 2 0 14 12 10 8 6 4 2 0 14 12 10 8 6 4 2 0 poly(dT)-poly(dT) LE poly(dG)-poly(dC) HE poly(dG)-poly(dC) HO poly(dG)-poly(dC) LE poly(dG)-poly(dC) LO poly(dA)-poly(dT) HE poly(dA)-poly(dT) HO poly(dA)-poly(dT) LO GCGC... HE GCGC... LE GCGC... LO GCGC... HO CGCG... HE CGCG... LE CGCG... LO CGCG... HO ATAT... LE ATAT... HE ATAT... LO ATAT... HO TATA... LE -- TATA... HE TATA... LO TATA... HO ACAC... HE CACA... HE GTGT... LE TGTG... LE-- ACAC... LE CACA... LE CTCT... HE TCTC... HE GAGA... HO AGAG... HO TCTC... LO CTCT... LO GAGA... HE AGAG... HE TCTC... LE CTCT... LE GAGA... LO AGAG... LO GTGT... HO TGTG... HO GAGA... LE AGAG... LE GTGT... HE TGTG... HE CACA... HO ACAC... HO GTGT... LO TGTG... LO ACAC... LO CACA... LO TCTC... HO CTCT... HO FIG. 23: (Color online) The exponent η of the power-law fits, k = k′ 0N −η, within TB II. through DNA molecules, Nature 398, 407 (1999). 11 D. Porath, A. Bezryadin, S. De Vries, and C. Dekker, Direct measurement of electrical transport through DNA molecules, Nature 403, 635 (2000). 12 A.J. Storm, J. van Noort, S. de Vries, and C. Dekker, In- sulating behavior for DNA molecules between nanoelec- trodes at the 100 nm length scale, Appl. Phys. Lett. 79, 3881 (2001). 13 K.-H. Yoo, D.H. Ha, J.-O. Lee, J.W. Park, J. Kim, J.J. Kim, H.-Y. Lee, T. Kawai, and H.Y. Choi, Electrical Conduction through Poly(dA)-Poly(dT) and Poly(dG)- Poly(dC) DNA Molecules, Phys. Rev. Lett. 87, 198102 (2001). 14 B. Xu, P. Zhang, X. Li, and N. Tao, Direct Conductance Measurement of Single DNA Molecules in Aqueous Solu- tion, Nano Lett. 4, 1105 (2004). 15 H. Cohen, C. Nogues, R. Naaman, and D. Porath, Direct measurement of electrical transport through single DNA molecules of complex sequence, Proc. Nat. Acad. Sci. 102, 11589 (2005). 16 G.P. Triberis, C. Simserides and V.C. Karavolas, Small polaron hopping transport along DNA molecules, J. Phys.: Condens. Matter 17 (2005) 2681. 17 C. Simserides, A systematic study of electron or hole trans- fer along DNA dimers, trimers and polymers, Chem. Phys. 440, 31 (2014). 18 K. Lambropoulos, M. Chatzieleftheriou, A. Morphis, K. Kaklamanis, M. Theodorakou, and C. Simserides, Unbi- ased charge oscillations in B-DNA: Monomer polymers and dimer polymers, Phys. Rev. E 92, 032725 (2015). 19 C.H. Wohlgamuth, M.A. McWilliams, and J.D. Slinker, DNA as a Molecular Wire: Distance and Sequence Depen- dence, Anal.Chem. 85, 8634 (2013). 20 F.D. Lewis, M.R. Wasielewski, Dynamics and efficiency of photoinduced charge transport in DNA: Toward the elu- sive molecular wire, Pure and Applied Chemistry 85, 1379 (2013). 21 R. Gutirrez, R. Caetano, P.B. Woiczikowski, T. Kubar, M. Elstner and G. Cuniberti, Structural fluctuations and quantum transport through DNA molecular wires: a com- bined molecular dynamics and model Hamiltonian ap- proach, New Journal of Physics 12, 023022 (2010). 22 K. Kawai, T. Majima, Chapter 44, Increasing the Hole Transfer Rate Through DNA by Chemical Modification, in Chemical Science of -Electron Systems, edited by T. Akasaka, A.O.S. Fukuzumi, H.K.Y. Aso, Springer Springer Tokyo Heidelberg New York Dordrecht London (2015). 23 Y.-J. Ye, L.-L. Shen, DFT Approach to Calculate Elec- tronic Transfer through a Segment of DNA Double Helix, J. Comp. Chem. 21 (2000) 1109. 24 Y.-J. Ye, Y. Jiang, Electronic Structures and Long-Range Electron Transfer Through DNA Molecules, Int. J. Quant. Chem. 78 (2000) 112. 25 R.N. Barnett, C.L. Cleveland, U. Landman, E. Boone, S. Kanvah, G.B. Schuster, Effect of Base Sequence and Hy- dration on the Electronic and Hole Transport Properties of Duplex DNA: Theory and Experiment, J. Phys. Chem. A 107 (2003) 3525. 26 E. Artacho, M. Machado, D. S´anchez-Portal, P. Ordej´on, and J.M. Soler, Electrons in dry DNA from density func- tional calculations, Mol. Phys. 101 (2003) 1587. 27 C. Adessi, S. Walch, and M.P. Anantram, Environment and structure influence on DNA conduction, Phys. Rev. B 67 (2003) 081405(R). 28 H. Mehrez and M. P. Anantram, Interbase electronic cou- pling for transport through DNA, Phys. Rev. B Vol. 71, 115405, 2005. 29 A.A. Voityuk, Electronic couplings and on-site energies for hole transfer in DNA: Systematic quantum mechani- cal/molecular dynamic study, J. Chem. Phys. 128 (2008) 115101. 30 G. Cuniberti, L. Craco, D. Porath, and C. Dekker, Backbone-induced semiconducting behavior in short DNA wires, Phys. Rev. B 65 (2002) 241314 (R). 31 S. Roche, D. Bicout, E. Maci´a, and E. Kats, Long Range Correlations in DNA: Scaling Properties and Charge Transfer Efficiency, Phys. Rev. Lett. 91 (2003) 228101. 32 S. Roche, Sequence Dependent DNA-Mediated Conduc- tion, Phys. Rev. Lett. 91 (2003) 108101. 33 F. Palmero, J.F.R. Archilla, D. Hennig, and F.R. Romero, Effect of base-pair inhomogeneities on charge transport along the DNA molecule, mediated by twist and radial polarons, New J. Phys. 6 (2004) 13. 34 H. Yamada, Localization of electronic states in chain mod- els based on real DNA sequence, Phys. Lett. A 332 (2004) 65. 35 V.M. Apalkov and T. Chakraborty, Electron dynamics in a DNA molecule, Phys. Rev. B 71 (2005) 033102. 36 D.K. Klotsa, R.A. Romer, and M.S. Turner, Electronic 20 Transport in DNA, Biophys. J. 89 (2005) 2187. 37 C.T. Shih, S. Roche, R.A. Romer, Point-Mutation Effects on Charge-Transport Properties of the Tumor-Suppressor Gene p53, Phys. Rev. Lett. 100 (2008) 018105. 38 Y.S. Joe, S.H. Lee, E.R. Hedin, Electron transport through asymmetric DNA molecules, Phys. Lett. A 374 (2010) 2367. 39 E.L. Albuquerque, U.L. Fulco, V.N. Freire, E.W.S. Caetano, M.L. Lyra, F.A.B.F. de Moura, DNA-based nanobiostructured devices: The role of quasiperiodicity and correlation effects, Physics Reports 535, 139 (2014). 40 G. Cuniberti, E. Maci´a, A. Rodr´ıguez, and R.A. Romer, Chapter: Tight-Binding Modeling of Charge Migration in DNA Devices, pp. 1-20 in T. Chakraborty (editor), Charge Migration in DNA Perspectives from Physics, Chemistry, and Biology, 2007, Springer-Verlag Berlin Heidelberg. 41 K. Lambropoulos, K. Kaklamanis, G. Georgiadis, and C. Simserides, THz and above THz electron or hole oscilla- tions in DNA dimers and trimers, Ann. Phys. (Berlin) 526, 249 (2014). 42 K. Lambropoulos, K. Kaklamanis, G. Georgiadis, M. Theodorakou, M. Chatzieleftheriou, M. Tassi, A. Morphis and C. Simserides, THz oscillations in DNA monomers, dimers and trimers, Proceedings of PIERS 2015 (Progress In Electromagnetics Research Symposium ), Prague, Czech Republic, 06-09 July, 2015. 43 L.G.D. Hawke, G. Kalosakas, and C. Simserides, Elec- tronic parameters for charge transfer along DNA, Eur. Phys. J. E 32 (2010) 291; ibid. Erratum to: Electronic parameters for charge transfer along DNA, 34 (2011) 118. 44 M.J.C. Gover, The eigenproblem of a tridiagonal 2- Toeplitz matrix, Linear Algebra and its Applications 197- 198 (1994) 63. 45 Said Kouachi, Eigenvalues and eigenvectors of tridiagonal matrices, Electronic Journal of Linear Algebra 15 (2006) 115. 46 R. Alvarez-Nodarse, J. Petronilho, N.R. Quintero, On some tridiagonal k-Toeplitz matrices: Algebraic and an- alytical aspects. Applications, Journal of Computational and Applied Mathematics 184 (2005) 518.
1110.0373
1
1110
2011-10-03T15:03:35
An excitable electronic circuit as a sensory neuron model
[ "physics.bio-ph", "q-bio.NC" ]
An electronic circuit device, inspired on the FitzHugh-Nagumo model of neuronal excitability, was constructed and shown to operate with characteristics compatible with those of biological sensory neurons. The nonlinear dynamical model of the electronics quantitatively reproduces the experimental observations on the circuit, including the Hopf bifurcation at the onset of tonic spiking. Moreover, we have implemented an analog noise generator as a source to study the variability of the spike trains. When the circuit is in the excitable regime, coherence resonance is observed. At sufficiently low noise intensity the spike trains have Poisson statistics, as in many biological neurons. The transfer function of the stochastic spike trains has a dynamic range of 6 dB, close to experimental values for real olfactory receptor neurons.
physics.bio-ph
physics
October 24, 2018 3:50 medeiros-ijbc-2011 International Journal of Bifurcation and Chaos c(cid:13) World Scientific Publishing Company 1 1 0 2 t c O 3 ] h p - o i b . s c i s y h p [ 1 v 3 7 3 0 . 0 1 1 1 : v i X r a An excitable electronic circuit as a sensory neuron model Bruno N. S. Medeiros1, Victor Minces2, Gabriel B. Mindlin3, Mauro Copelli1, Jos´e R. Rios Leite1 1Departamento de F´ısica, Universidade Federal de Pernambuco, 50670-901, Recife, PE, Brazil. 2Department of Cognitive Neuroscience, University of California San Diego, 92093-0515, La Jolla, CA, USA 3Departamento de F´ısica, FCEN, Universidad de Buenos Aires, Ciudad Universitaria, Pab. I, 1428, Buenos Aires, Argentina Received (to be inserted by publisher) An electronic circuit device, inspired on the FitzHugh-Nagumo model of neuronal excitability, was constructed and shown to operate with characteristics compatible with those of biological sensory neurons. The nonlinear dynamical model of the electronics quantitatively reproduces the experimental observations on the circuit, including the Hopf bifurcation at the onset of tonic spiking. Moreover, we have implemented an analog noise generator as a source to study the variability of the spike trains. When the circuit is in the excitable regime, coherence resonance is observed. At sufficiently low noise intensity the spike trains have Poisson statistics, as in many biological neurons. The transfer function of the stochastic spike trains has a dynamic range of 6 dB, close to experimental values for real olfactory receptor neurons. Keywords: Electronic circuit, Excitable element, Coherence resonance, Dynamic range. 1. Introduction Ever since the pioneering work of Hodgkin & Huxley [1952], the biophysical mechanisms underlying the generation and propagation of action potentials (spikes) in neurons have been described with increasing detail, ranging from the discovery of new types of ion channels to the study of intracellular calcium dynamics [Hille, 2001]. No matter how interesting, these new findings have helped little in our understanding of collective neuronal phenomena, which remain a daunting task in face of the interplay among high- dimensionality, noise and nonlinearity (see e.g. Chialvo [2010] for a recent review). The challenge should nonetheless be faced: the solution of issues at the frontiers of current-day neuroscience, like e.g. grandmother cell [Barlow, 1972] versus population coding [Young & Yamane, 1992], or firing rate versus spike-time coding [Rieke et al., 1999] will likely be grounded on our success in this endeavor. In fact, theoretical progress in this front has been achieved in recent years with very simple models. One such example is the proposed solution for the century-old problem of the origin of psychophysical response curves [Copelli et al., 2002; Kinouchi & Copelli, 2006]. Steven's psychophysical law states that the psychological perception F of a physical stimulus (e.g. light, or odorant) of intensity h is a power law F ∝ hs, with experimental values of the Stevens exponent s fluctuating around s ≃ 0.5. Compared to a linear response, psychophysical nonlinear responses have at least one evolutionarily favorable property: 1 October 24, 2018 3:50 medeiros-ijbc-2011 2 Medeiros et al. R3=1MΩ R4=10k Ω Vin C = 1nF or 50pF R1=1k Ω V_ V+ Va Vb R5=10k Ω Vout TL071 R2=10k Ω Fig. 1. Excitable electronic circuit. Va and Vb = −Va are the operational amplifier supply voltages. Vin is an input voltage, corresponding to an external stimulus. We describe the circuit as a two-dimensional dynamical system on the variables V− and Vout (see Eqs. (2)). they amplify weaker stimulus, i. e. they have a larger dynamic range. But how do the Stevens exponents arise in the nervous system? At first, this question seems puzzling because single neurons typically have small dynamic ranges [Rospars et al., 2000]. A theoretical solution recently proposed involves a collective phenomenon: excitable waves are generated by the incoming stimuli and propagate "laterally" among excitable neu- rons, thereby amplifying the system response (in comparison to what would be observed in the absence of the coupling). Interestingly, this amplification mechanism is self-limited: under intense stimulation, for instance, a large number of excitable waves can be created, but owing to refractoriness they annihilate upon collision. The enhancement of dynamic range in this model is therefore governed by the low-stimulus amplification [Copelli et al., 2002; Kinouchi & Copelli, 2006]. Robustness of these results has been tested at different modeling levels [Copelli et al., 2005; Ribeiro & Copelli, 2008; Assis & Copelli, 2008; Publio et al., 2009], showing that the degree of biophysical realism in the model of each neuron is less relevant to the global dynamics than the topology of the network [Copelli & Kinouchi, 2005; Copelli & Campos, 2007; Ribeiro & Copelli, 2008; Assis & Copelli, 2008; Gollo et al., 2009]. This phenomenon has also been studied analytically [Furtado & Copelli, 2006; Larremore et al., 2011] and was recently confirmed experimentally in cortical slices [Shew et al., 2009]. The appeal of a sensory system with large dynamic range based on a network of simple excitable units, each with small dynamic range, goes beyond basic research in neuroscience. The idea could be reversed, leading to biologically inspired artificial sensors, which have been used in a variety of scenarios (see e.g. de Souza et al. [1999]). There are several electronic circuits reported in the literature which have been designed to present a neuron-like dynamical response. The rationale behind those efforts was to dynamically interact with biological neurons rather than stimulating them using response independent current commands. In this way, electronic circuits which analogically integrated the Hindmarsh and Rose equations [Szucs et al., 2000] were coupled to the neurons of a preparation of lobster pyloric CPG neurons. This allowed to show that regularity could emerge as a collective dynamical property of units which individually presented complex dynamics. In another set of experiments, electronic neurons interacting with a biological preparation were used to unveil which dynamical properties of a neural network depend on the bifurcation leading to excitation for the units, rather than on the details of the neural dynamics. To carry out this program, a standard form for class I excitable dynamics was analogically integrated with a circuit, which was used to replace a neuron in a midbody ganglion of the leech Hirudo medicinalis [Aliaga et al., 2003]. The responses under the stimulation of both the natural preparation and the one with a replaced neuron were found to be similar. Beyond the possibility of interacting with neurons through a dynamically sensible way, these efforts provide empirical support to the program of studying neural processes through simple and relatively low dimensional dynamical systems. Depending on the question under study, it might be desirable to be able to October 24, 2018 3:50 medeiros-ijbc-2011 An excitable electronic circuit as a sensory neuron model 3 establish a closer link between the device and a neuron. In this spirit, a device implementing a conductance model was recently proposed [Sitt & Aliaga, 2007]. These circuits, however, have two limitations for our purposes. First, they are still too complex to be replicated in large scale. Second, they do not have a controllable noise source to produce stochastic spike trains, a feature that is common to the both models [Copelli et al., 2002] and real neurons [Dayan & Abott, 2001; Mainen & Sejnowski, 1995; Petracchi et al., 1995]. The present work is a first step in this direction. We propose an excitable electronic circuit which can serve as a building block of an electronic sensor. The advantages of its extreme simplicity are twofold: it allows for scalability and, at the same time, simple mathematical modeling. The paper is organized as follows. In section 2, we describe the electronic circuit and the equations that model its dynamics. In section 3, we introduce noise from a simple analog noise generator at the input of the excitable circuit and study the statistical properties of the resulting spike trains and show that it can exhibit Poisson statistics as well as coherence resonance, as expected. In section 4 we evaluate the dynamic range of the excitable circuit and show that it is comparable to that of single sensor neurons. 2. Dynamic model The circuit we propose is shown in Fig. 1. It is composed of five resistors, one capacitor and one operational amplifier. The voltage Vin corresponds to an external stimulus, which can be e.g. a constant or the sum of DC and noise voltages. In our electronic neuron, the operational amplifier behaves as a simple comparator circuit, for which we use the following nonlinear model: dVout dt = S sign [Vb − Vout + (Va − Vb)Θ(V+ − V−)] , (1) where Θ is the Heaviside function and S is the op-amp slew rate (whose datasheet value for the simple TL071 in the circuit is S = 16 V/s). As usual, symmetric supply voltages Vb = −Va were used. Assuming R3 ≫ R4, R5 and applying Kirchhoff's laws, we arrive at a two-dimensional dynamic model on the variables Vout and V−: dVout dt dV− dt = = Vc ε sign [Vb − Vout + (Va − Vb)Θ(αVout − V−)] , 1 R3C(cid:2)βVout + γVin − V−(cid:3) . (2a) (2b) where α ≡ R1/(R1 + R2), β ≡ R4/(R4 + R5) and γ ≡ R5/(R4 + R5). Vc = 10 V is a characteristic voltage of the same order of magnitude of the supply voltages, and we have defined ε ≡ Vc/S as a characteristic (short) time scale. To avoid the possibility that the system (2) has more than one fixed point, we require β > α. In terms of the variables v ≡ w ≡ , Vout Vc V− Vc , the equations can be rewritten in dimensionless form v = sign(cid:18)b − v + (a − b) 1 + e−(αv−w)/x0(cid:19) , w = φ [βv + γj − w] , where we defined the dimensionless groups: τ ≡ t ε ; a = Va Vc ; b = Vb Vc ; φ = ε R3C ; j = Vin Vc , (3a) (3b) (4a) (4b) (5) October 24, 2018 3:50 medeiros-ijbc-2011 4 Medeiros et al. a) a) 0.1 0.1 0.05 0.05 w w 0 0 -0.05 -0.05 -0.1 -0.1 b) ) V 1.5 1 0.5 ( - 0 V -0.5 -1 -1.5 c) 1.5 1 ) z H k ( f 0.5 0 -1 -0.5 0 0.5 1 -1 -0.5 0 0.5 1 -15-10 -5 0 5 10 15 -10 -5 0 5 10 d) v v 15 10 ) V ( t u o V 5 0 -5 -10 -15 Vout (V) e) 10 Vin (V) ) V ( - V ) V ( m V 5 0 -5 -10 0 1 2 3 4 0 1 2 3 4 time (ms) time (ms) a) Nullclines of system (4) for a = 1, b = 1, α = 0.0909, β = 0.5, γ = 0.5, j = 0, φ = 0.01 and x0 = 9 × 10−3: Fig. 2. solid black line for the v = 0 nullcline and dashed black line for the w = 0 nullcline. The fixed point is unstable and the trajectories are attracted to a limit cycle (red solid line). b) Experimental limit cycle (black dots) and numerical integration of the model (red solid line) for x0 = 1 × 10−5, Va = 10 V, Vb = −10 V, Vin = −6 V and φ = 5 × 10−4 (other parameters are the same as in (a)). c) Experimental frequency response f to the external DC stimulus Vin (black dots) and the same for the numerical integration of the model (red line). d) Comparison between experimental time series of Vout and V− (black circles and triangles, respectively) with numerical integration of the model (red and green lines, respectively). e) Experimental (black dots) and numerical (red line) spike trains obtained from the analog subtraction Vm of the dynamical variables (see text for details). and replaced Θ by the continuous function Θ(x; x0) = 1 1 + e−x/x0 (6) for the purpose of numerical integration and derivation (see below). Note that Θ → Θ as x0 → 0. The constant φ ≪ 1 sets the ratio between the fast and slow time scale as in the FitzHugh-Nagumo model, so that R3C ultimately controls the overall time scale of the problem. As shown in Fig. 2a (black lines), the nullclines v = 0 and w = 0 of Eqs. (4) resemble those of the FitzHugh-Nagumo model for neuronal excitability, with one fast (v or Vout) and one slow (w or V−) variable. In the limit x0 → 0, the cubic-like v = 0 nullcline becomes piecewise linear. When the fixed point sits at its outer branches, it is stable. It loses stability in a Hopf bifurcation as the w nullcline crosses the v nullcline at its central branch, so trajectories are attracted to a limit cycle (red line) with nonzero frequency f (i.e. f changes discontinuously at the bifurcation). Below the Hopf bifurcation, the circuit is said to be type-II excitable [Rinzel & Ermentrout, 1998]. There is good quantitative agreement between experimental data from the circuit and the numerical integration, as can be seen in Fig. 2b, c, d and e. Note that through an analog subtraction Vm ≡ 1.5V− − 0.67Vout (see also Fig. 3b) the circuit exhibits the spikes typical of neuronal membrane potentials (Fig. 2e). We emphasize that in Fig. 2 experimental and numerical data agree without any fitting parameter, as long as x0 is sufficiently small (. 10−4). 3. Noise addition and coherence resonance So far we have discussed the response of the excitable circuit under DC stimulation. Biological neurons, however, can show highly variable responses, even when subjected to a presumably constant stimulus. October 24, 2018 3:50 medeiros-ijbc-2011 An excitable electronic circuit as a sensory neuron model 5 a) 10k 12V 100nF TL074CN KN2222 100k 1k R A1 0−50k R A2 100nF 1k R A3 10k−100k R A4 TL074CN Vnoise 1 Vin Analog addition 2 Excitable circuit Vout 3 Buffer V_ V_ 4 Analog subtraction Vm _ 1.5 V_ 0.67 Vout b) VDC Vnoise Fig. 3. a) An analog noise generator based on the amplification of transistors thermal noise. Noise amplification is given by A = [(RA1 + RA2)/RA1](RA4/RA3). b) Block diagram of the circuit used to verify the excitability of the circuit pre- sented in Fig. 1. Analog addition and subtraction (see text for details) are performed with standard TL074 op-amp opera- tions [Senturia & Wedlock, 1981]. experimental numerical tp b) c) ) V ( m V ) V ( m V 8 4 0 -4 -8 8 4 0 -4 -8 -2 -2 -1 -1 0 0 v v 1 1 2 2 0 0.1 Poisson 0.2 0.3 time (s) 0.4 0.5 a) a) 0.1 0.1 w w 0 0 -0.1 -0.1 d) 100 p R 101 102 noise amplification A 103 104 Fig. 4. a) Numerical phase plane trajectory (red line) due to noise excitation. Without noise the system would stay in a resting state at the fixed point (white dot). Experimental (b) and numerical (c) spike train series are shown when the system is in the excitable state (stable fixed point as shown in (a)). d) Experimental coherence resonance curve for C = 50 pF, Va = 12 V = −Vb and VDC = −7.826 V (see Fig. 1). Each point corresponds to an average over 10 s time series. Examples range from highly variable responses olfactory receptor neurons (ORNs) to presentation of iden- tical puffs of odorants [Rospars et al., 2000], to cortical cells stimulated with a constant current via an intracellular electrode [Mainen & Sejnowski, 1995]. In an attempt to endow our excitable circuits with the variability in the spike trains observed in biological neurons, we propose the simple analog noise generator shown in Fig. 3a. Once more, its simplicity allows one to attach independent noise generators to each excitable circuit when connecting them in a network. The circuit in Fig. 3a provides a two-stage amplification control via two operational amplifiers to the thermal noise produced by the KN2222 transistors. Its output voltage Vnoise is approximately a Gaussian white noise voltage with a cutoff frequency around 1 kHz. To obtain variable spike trains, the stimulus Vin consists in the analog addition of VDC and Vnoise (see October 24, 2018 3:50 medeiros-ijbc-2011 6 Medeiros et al. blocks 1 and 2 in Fig. 3b). In the model, this corresponds to replacing Eq. (4b) with w = φ [βv + γj + Dξ(t) − w] , (7) where D grows linearly with the gain in the noise amplification A (which in turn is controlled by the variable resistors shown in Fig. 3). Setting VDC below the Hopf bifurcation, the circuit sits at a stable fixed point at the right branch of the v = 0 nullcline, from which it eventually departs owing to noise (Fig. 4a). This generates spike trains with variable interspike intervals tp, as shown in Fig. 4b and c. We now show that the interplay between noise and excitability behaves as expected in our simple circuits. Pikovsky & Kurths [1997] have shown that the coherence the spike train of the FitzHugh-Nagumo model peaks at an intermediate noise value, in a phenomenon which has been called "coherence resonance". In other words, the normalized standard deviation Rp ≡ qht2 pi − htpi2 htpi (8) should have a minimum as a function of the noise intensity. This is precisely what we observe in our circuit when VDC (= −7.826 V) is close to the Hopf bifurcation (VHopf = −7.82 V), as displayed in Fig. 4d. Note that Rp close to zero means that the time series is approximately periodic. For small noise amplitudes (Vnoise ∼ 50 mV, or A ∼ O(1) in Fig. 4d), spikes are sparse and Rp approaches unity. This suggests a Poisson process in which the interspike interval distribution approaches an exponential P (tp) = re−rtp, (9) where r is time rate constant. This Poisson limit is interesting because it is observed in different neuronal preparations [Dayan & Abott, 2001; Petracchi et al., 1995], so we performed a detailed statistical analysis of the small Vnoise regime. a) 100 10-1 ) t ( D 10-2 10-3 10-4 c) 160 120 2 σ 80 40 0 fit e-rt r=10.0+-0.3 s-1 160 80 0 0 0.2 isi (s) 0.4 i s e k p s f o # 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 t (s) σ2 fit rT r=8.94+-0.03 s-1 b) 100 > n < 80 60 40 20 0 d) 160 120 2 σ 80 <n> fit rT r=9.060+-0.006 s-1 1 2 3 4 5 6 7 8 9 10 time window: T (s) fit k<n> k=0.99+-0.03 1 2 3 4 5 6 7 8 9 10 11 time window: T (s) 40 0 0 20 40 60 80 100 <n> Fig. 5. Spike train statistics of a 100 s duration series from the circuit of Fig. 3b. a) Interspike interval (isi) accumulated distribution in log-linear scale. Inset: corresponding histogram of isi. The dashed line corresponds to an exponential fit of a Poisson distribution with mean firing rate r = 10.0(3) s−1. The deviation from the Poisson distribution for small isi is due to the refractoriness of the excitable circuit. In the following graphs we have divided the series in time windows of duration T . The mean number of spikes hni (b) and the variance σ2 (c) are shown as functions of T . In (d) we have σ2 as a function of hni. The dashed lines are fits of σ2 = hni = rT according to the Poisson distribution. October 24, 2018 3:50 medeiros-ijbc-2011 An excitable electronic circuit as a sensory neuron model 7 In Fig. 5a the statistics of a 100 s experimental time series was compared to the accumulated distri- bution D(t) ≡ Z ∞ t re−rtpdtp = e−rt, (10) showing good agreement for a fitted rate r ≃ 10.0(3) s−1. To check for consistency, we divided the time series in small time windows of size T and sampled the number n of of spikes per window. In a Poisson n ≡ hn2i − hni2 = rT which are confirmed in Fig. 5b process one has the linear relationships hni = rT , σ2 and c. The unit slope in the σ2 n versus hni plot is also verified (see Fig. 5d). These results show that our circuit can be used to mimic not only deterministic dynamics, but also simple statistical properties which appear in biological neurons. 4. Dynamic range In this section we study the response of our excitable system to varying input voltage VDC, considering the noise amplitude Vnoise constant. Although in real neurons the background noise may have a dependence on the stimulus, it is a fair approximation to treat the noise amplitude as constant and focus on the dependence on input signal as a control parameter of the dynamics. In what follows, the response of the circuit is defined as the mean firing rate F measured over a fixed time interval Tm. This so-called "rate coding" is also a longstanding approximation [Adrian, 1926], which seems to fit data in several cases [Koch, 1999; Arbib, 2002]. For fixed Tm and noise amplification A, the response F of our circuit is an increasing function of the stimulus VDC because larger values of VDC amounts to increased excitability, lowering the "effective thresh- old" to noise-induced spike generation (there is no real threshold in type-II excitable neurons [Izhikevich, 2007]). Conversely, for fixed VDC, the response F also increases with increasing noise intensity A. These results are shown in Fig. 6a, where we plot (for different noise intensities) the responses F (VDC) of our excitable circuit with a 1 nF capacitor. This choice sets the time scale of the neuron in the millisecond range (i.e. that of biological neurons). Note that in the absence of noise (A = 0) the response is null up to Hopf bifurcation A = 2000 A = 1500 A = 1000 A = 500 A = 0 a) ) s / 1 ( F 300 250 200 150 100 50 0 b) ) s / 1 ( F 5000 4000 3000 2000 1000 0 Hopf bifurcation A = 2000 A = 1500 A = 1000 A = 500 A = 0 -8.05 -8 -7.95 -7.9 VDC (V) c) ) s / 1 ( F A = 1000 A = 0 200 150 100 V0.1 VHopf V0.9 50 0 0 ∆ = 6.62 dB 0.02 0.04 VDC-V0 (V) 0.06 -7.85 -7.8 -9.4 -9.2 -9 -8.8 -8.6 -8.4 -8.2 -8 -7.8 -7.6 VDC (V) C = 1 nF C = 50 pF d) 10 ) B d ( ∆ 9 8 7 6 5 4 3 2 0.08 0 500 1000 1500 2000 2500 A Fig. 6. Experimental response curves F (VDC) measured at different values of the noise amplification A. Supply voltages Va = 12 V = −Vb. a) C = 1 nF (φ = 5 × 10−4 and Tm = 10 s). b) C = 50 pF (φ = 0.01 and Tm = 0.2 s). c) Response curve for C = 1 nF (A=1000), and relevant parameter for calculating the dynamic range. d) Dynamic range as function of noise amplification for C = 50 pF (black squares) and C = 1 nF (white circles). October 24, 2018 3:50 medeiros-ijbc-2011 8 REFERENCES the Hopf bifurcation (so the lowest curve in Fig. 6a is similar to Fig. 2c). Results in Fig. 6b correspond to a circuit with a 50 pF capacitor. This single change renders a much faster circuit, now operating in the microsecond range, but with its dynamical features otherwise preserved. This has potential applications, because a faster circuit requires shorter measurement intervals Tm (=0.2 s in our example) for a reliable estimation of the firing rate. Given a response curve, we can calculate its dynamic range, which roughly speaking corresponds to the range of stimulus intensity that the firing rate can "appropriately code". Measured in decibels, this is arbitrarily defined as [Rospars et al., 2000; Copelli & Kinouchi, 2005] ∆ ≡ 10 log10(cid:18) V ∗ 0.9 V ∗ 0.1(cid:19) , (11) where V ∗ x ≡ Vx − V0 are measured relative to the voltage V0 at which the response becomes non-zero and F (Vx) = xFmax (0 ≤ x ≤ 1) , (12) where Fmax is the firing rate at the Hopf bifurcation. In words (see Fig. 6c), ∆ measures the range of stimulus VDC which are neither too small (VDC < V0.1) to go undetected nor too close (VDC > V0.9) to the autonomous oscillations that emerge at VHopf . As shown in Fig. 6d, the dynamic range is a rather robust feature of our excitable circuit: it changes little as the noise intensity is varied, regardless of the time scale at which it operates. In both cases, ∆ ≃ 6 dB, which is closer to the values obtained experimentally (∆ ≃ 10 dB for olfac- tory sensory neurons [Rospars et al., 2000], ∆ ≃ 14 dB for retinal ganglion cells [Deans et al., 2002; Furtado & Copelli, 2006]) than results obtained theoretically for discrete models of excitable elements (∆ ≃ 14 dB in [Furtado & Copelli, 2006] and ∆ ≃ 19 dB in [Assis & Copelli, 2008]). 5. Concluding remarks In summary, we have presented an excitable electronic circuit whose simplicity allows for scalability and accurate mathematical modeling. Its dynamical equations lead to time series which quantitatively reproduce experimental results without fitting parameters. In addition, we have shown that the introduction of noise from a simple analog noise generator at the input of the circuit produces variable spike trains. The statistics of the interspike intervals is shown to exhibit coherence resonance. Furthermore, by analyzing long time series under low noise intensity, the spike trains were shown to behave as a Poisson process, like some biological neurons. In the excitable regime, with fixed noise amplitude, the firing rate response of the system to a VDC input -- the stimulus -- was shown to have a dynamic range of about 6 dB, which is also comparable to some biological sensory neurons. Together with its scalability, these properties render the system a potential building block for artificial sensors based on collective properties of excitable media. 6. Acknowledgments BNSM, MC and JRRL acknowledge financial support from Brazilian agencies CNPq, FACEPE, CAPES and special programs PRONEX, PRONEM and INCEMAQ. GBM acknowledges support from NIH. It is a pleasure to thank Hugo L. D. S. Cavalcante for enlightening discussions during the preparation of this work, as well as Marcos Nascimento for technical support. References Adrian, E. D. [1926] "The impulses produced by sensory nerve endings: Part I," J. Physiol. (London) 61, 49 -- 72. Aliaga, J., Busca, N., Minces, V., Mindlin, G. B., Pando, B., Salles, A. & Sczcupak, L. [2003] "Electronic neuron within a ganglion of a leech (Hirudo medicinalis)," Phys. Rev. E 67, 061915. Arbib, M. A. [2002] The handbook of brain theory and neural networks (The MIT Press). October 24, 2018 3:50 medeiros-ijbc-2011 REFERENCES 9 Assis, V. R. V. & Copelli, M. [2008] "Dynamic range of hypercubic stochastic excitable media," Phys. Rev. E 77, 011923. Barlow, H. B. [1972] "Single units and sensation: A neuron doctrine for perceptual psychology," Perception 1, 371 -- 394. Chialvo, D. R. [2010] "Emergent complex neural dynamics," Nat. Phys. 6, 744 -- 750. Copelli, M. & Campos, P. R. A. [2007] "Excitable scale-free networks," Eur. Phys. J. B 56, 273 -- 278. Copelli, M. & Kinouchi, O. [2005] "Intensity coding in two-dimensional excitable neural networks," Physica A 349, 431 -- 442. Copelli, M., Oliveira, R. F., Roque, A. C. & Kinouchi, O. [2005] "Signal compression in the sensory periphery," Neurocomputing 65-66, 691 -- 696. Copelli, M., Roque, A. C., Oliveira, R. F. & Kinouchi, O. [2002] "Physics of Psychophysics: Stevens and Weber-Fechner laws are transfer functions of excitable media," Phys. Rev. E 65, 060901. Dayan, P. & Abott, L. F. [2001] Theoretical Neuroscience: Computational and Mathematical Modeling of Neural Systems (The MIT Press). de Souza, J. E. G., Neto, B. B., dos Santos, F. L., de Melo, C. P., Santos, M. S. & Ludermir, T. B. [1999] "Polypyrrole based aroma sensor," Synth. Met. 102, 1296 -- 1299. Deans, M. R., Volgyi, B., Goodenough, D. A., Bloomfield, S. A. & Paul, D. L. [2002] "Connexin36 is essential for transmission of rod-mediated visual signals in the mammalian retina," Neuron 36, 703 -- 712. Furtado, L. S. & Copelli, M. [2006] "Response of electrically coupled spiking neurons: a cellular automaton approach," Phys. Rev. E 73, 011907. Gollo, L. L., Kinouchi, O. & Copelli, M. [2009] "Active dendrites enhance neuronal dynamic range," PLoS Comput. Biol. 5, e1000402. Hille, B. [2001] Ion Channels of Excitable Membranes, 3rd ed. (Sinauer Associates (Sunderland)). Hodgkin, A. L. & Huxley, A. F. [1952] "A quantitave description of membrane current ands its application to conduction an excitation in nerve," J. Physiol. 117, 500 -- 544. Izhikevich, E. M. [2007] Dynamical Systems in Neuroscience: The Geometry of Excitability and Bursting (The MIT Press, London). Kinouchi, O. & Copelli, M. [2006] "Optimal dynamical range of excitable networks at criticality," Nature Physics 2, 348 -- 351. Koch, C. [1999] Biophysics of Computation (Oxford University Press, New York). Larremore, D. B., Shew, W. L. & Restrepo, J. G. [2011] "Predicting criticality and dynamic range in complex networks: Effects of topology," Phys. Rev. Lett. 106, 058101. Mainen, Z. & Sejnowski, T. [1995] "Reliability of spike timing in neocortical neurons," Science 268, 1503 -- 1506. Petracchi, D., Barbi, M., Chillemi, S., Pantazelou, E., Pierson, D., Dames, C., Wilkens, L. & Moss, F. [1995] "A test for a biological signal encoded by noise," Int. J. Bifurcat. Chaos 5, 89 -- 100. Pikovsky, A. S. & Kurths, J. [1997] "Coherence resonance in a noise-driven excitable system," Phys. Rev. Lett. 78, 775 -- 778. Publio, R., Oliveira, R. F. & Roque, A. C. [2009] "A computational study on the role of gap junctions and rod Ih conductance in the enhancement of the dynamic range of the retina," PLoS ONE 4, e6970. Ribeiro, T. L. & Copelli, M. [2008] "Deterministic excitable media under Poisson drive: Power law re- sponses, spiral waves and dynamic range," Phys. Rev. E 77, 051911. Rieke, F., Warland, D., de Ruyter van Steveninck, R. & Bialek, W. [1999] Spikes: Exploring the Neural Code (MIT Press, Cambridge, MA). Rinzel, J. & Ermentrout, B. [1998] "Analysis of neural excitability and oscillations," Methods in Neuronal Modeling: From Ions to Networks, eds. Koch, C. & Segev, I., 2nd ed. (MIT Press), pp. 251 -- 292. Rospars, J.-P., L´ansk´y, P., Duchamp-Viret, P. & Duchamp, A. [2000] "Spiking frequency versus odorant concentration in olfactory receptor neurons," BioSystems 58, 133 -- 141. Senturia, S. D. & Wedlock, B. D. [1981] Electronic Circuits and Applications (John Wiley). Shew, W., Yang, H., Petermann, T., Roy, R. & Plenz, D. [2009] "Neuronal avalanches imply maximum dynamic range in cortical networks at criticality," J. Neurosci. 29, 15595 -- 15600. October 24, 2018 3:50 medeiros-ijbc-2011 10 REFERENCES Sitt, J. D. & Aliaga, J. [2007] "Versatile biologically inspired electronic neuron," Phys. Rev. E 76, 051919. Szucs, A., Varona, P., Volkovskii, A. R., Abarbanel, H. D. I., Rabinovich, M. & Selverston, A. I. [2000] "Interacting biological and electronic neurons generate realistic oscillatory rhythms," Neuroreport 11, 563 -- 569. Young, M. & Yamane, S. [1992] "Sparse population coding of faces in the inferotemporal cortex," Science 256, 1327 -- 1331.
1007.4529
1
1007
2010-07-26T18:25:34
Amino Acids in Comets and Meteorites: Stability under Gamma Radiation and Preservation of Chirality
[ "physics.bio-ph", "astro-ph.EP" ]
Amino acids in solar system bodies may have played a key role in the chemistry that led to the origin of life on Earth. We present laboratory studies testing the stability of amino acids against gamma radiation photolysis. All the 20 chiral amino acids in the levo form used in the proteins of the current terrestrial biochemistry have been irradiated in the solid state with gamma radiation to a dose of 3.2 MGy which is the dose equivalent to that derived by radionuclide decay in comets and asteroids in 1.05x109 years. For each amino acid the radiolysis degree and the radioracemization degree was measured by differential scanning calorimetry (DSC) and by optical rotatory dispersion (ORD) spectroscopy. From these measurements a radiolysis rate constant kdsc and a radioracemization rate constant krac have been determined for each amino acid and extrapolated to a dose of 14 MGy which corresponds to the expected total dose delivered by the natural radionuclides decay to all the organic molecules present in comets and asteroids in 4.6x109 years, the age of the Solar System. It is shown that all the amino acids studied can survive a radiation dose of 14 MGy in significant quantity although part of them are lost in radiolytic processes. Similarly, also the radioracemization process accompanying the radiolysis does not extinguish the chirality. The knowledge of the radiolysis and radioracemization rate constants may permit the calculation of the original concentration of the amino acids at the times of the formation of the Solar System starting from the concentration found today in carbonaceous chondrites. For some amino acids the concentration in the presolar nebula could have been up to 6 times higher than currently observed in meteorites.
physics.bio-ph
physics
AMINO ACIDS IN COMETS AND METEORITES: STABILITY UNDER GAMMA RADIATION AND PRESERVATION OF CHIRALITY Susana Iglesias-Groth1,2, Franco Cataldo3,4, Ornella Ursini5, Arturo Manchado1,2,6 1Instituto de Astrofísica de Canarias (IAC), Via Lactea s/n, E-38200 La Laguna, Tenerife, Spain. 2Departamento de Astrofísica de la Universidad de La Laguna, Spain 3Istituto Nazionale di Astrofisica. Osservatorio Astrofisico di Catania, Via S. Sofia 78, 95123 Catania, Italy. 4Actinium Chemical Research, Via Casilina 1626/A, 00133 Rome, Italy. 5Istituto di Metodologie Chimiche, CNR, Via Salaria Km 29,300, 00016 Monterotondo Stazione, Rome, Italy. 6CSIC, Spain. Abstract Amino acids in solar system bodies may have played a key role in the chemistry that led to the origin of life on Earth. We present laboratory studies testing the stability of amino acids against γ radiation photolysis. All the 20 chiral amino acids in the levo form used in the proteins of the current terrestrial biochemistry have been irradiated in the solid state with γ radiation to a dose of 3.2 MGy which is the dose equivalent to that derived by radionuclide decay in comets and asteroids in 1.05x109 years. For each amino acid the radiolysis degree and the radioracemization degree was measured by differential scanning calorimetry (DSC) and by optical rotatory dispersion (ORD) spectroscopy. From these measurements a radiolysis rate constant kdsc and a radioracemization rate constant krac have been determined for each amino acid and extrapolated to a dose of 14 MGy which corresponds to the expected total dose delivered by the natural radionuclides decay to all the organic molecules present in comets and asteroids in 4.6x109 years, the age of the Solar System. It is shown that all the amino acids studied can survive a radiation dose of 14 MGy in significant quantity although part of them are lost in radiolytic processes. Similarly, also the radioracemization process accompanying the radiolysis does not extinguish the chirality. The knowledge of the radiolysis and radioracemization rate constants may permit the calculation of the original concentration of the amino acids at the times of the formation of the Solar System starting from the concentration found today in carbonaceous chondrites. For some amino acids the concentration in the presolar nebula could have been up to 6 times higher than currently observed in meteorites. It is also expected the preservation of an original chiral excess. This study adds experimental support to the suggestion that amino acids were formed in the interstellar medium and in chiral excess and then were incorporated in comets and asteroids at the epoch of the Solar System formation. Key words: astrobiology , astrochemistry, comets, meteorites, methods:laboratory 1. Introduction One of the breakthroughs in astrochemistry and astrobiology occurred in recent years was the unequivocal discovery of chirality in carbonaceous chondrites (CC), the rather rare meteorites which contain a relatively high level of macromolecular carbon commonly known as kerogen (Cronin and Pizzarello, 1997; Pizzarello and Cronin 2000; Pizzarello et al. 2008). More in detail, the enantiomeric excess was measured in the amino acids fraction extracted from the CC. The absolute certainty that the measured enantiomeric excess does not derive from contaminations is due to the fact that not only common α-amino acids used in the biochemistry of living organisms have been found in the extracts from CC but even really uncommon β-amino acids (Pizzarello and Cronin, 2000; Pizzarello et al. 2008) and diaminoacids (Meierhenrich et al. 2004; Meierhenrich, 2008). This fact excludes any contamination from terrestrial microorganisms and confirms the abiotic origin of the observed chiral excess. Furthermore also the isotopic signature confirms the non-terrestrial origin of the organic molecules extracted from the CC (Engel and Macko, 1997). As a further proof of the abiotic origin of the amino acids present in meteorites it is reported that the distribution of the abundances of the different amino acids follows that observed in the laboratory abiotic synthesis according to the Miller (1953,1955 and 2000) or the Sagan and Khare (1971) derived reactions. The abiotic origin of chirality is hence an experimental fact, and chirality is a key property of the organic molecules involved in the biochemistry of life. Without chirality there is no life because all the chemical reactions occurring in living organism are asymmetric synthesis and cannot occur or cannot be as fast and as selective as they are thank to chirality of the molecules and macromolecules involved (Lough and Weiner, 2002; Cataldo, 2007a). There are several hypotheses on the mechanism which has led to the experimentally observed chiral excess. The most convincing mechanism involves the action of circularly polarized light (CPL) on racemic mixtures of molecules formed in the interstellar medium before the formation of the Solar System (Meierhenrich, 2008). Because of the slightly different light absorption cross section in a racemic mixture one enantiomer absorbs a slightly higher amount of CPL than the other enantiomer. A prolonged irradiation of a racemic mixture of molecules with CPL transform it into a scalemic mixture, hence with a chiral excess. The effect of UV-CPL has already been demonstrated experimentally in creating a chiral excess (Jorissen and Cerf, 2002; Meierhenrich et al. 2005; Meierhenrich 2008). Thus, the scalemic mixture of molecules once formed in the interstellar medium where then incorporated in the primitive Solar nebula and can be found still today in meteorites and presumably other bodies of the Solar System which have experienced only a limited processing, in other words where the matter was preserved as it was 4.6 x 109 year ago (Sephton 2002 ; Kwok 2009). The point is that there are already known sources of circularly polarized light in certain star-forming regions (Bailey et al. 1998) which have a clear implication in the induction of chirality in the racemic mixture of molecules formed with abiotic processes in the interstellar medium. To further underline the complexity of the kerogen in the CC, it should be remembered that only 10-20% of the carbonaceous matter is water or solvent extractable, the remaining fraction is composed by a complex mixture of high molecular weight crosslinked organic matter resembling for certain instances the coal of terrestrial origin. Refined analytical techniques have been used to study also this insoluble fraction revealing that it is a really complex mixture of thousands of different molecular species (Schmitt-Kopplin 2010, Sephton, 2002), a complexity comparable or even more impressive than that of terrestrial coal. The top down approach in the search of the abiotic synthesis of amino acids and the relative chiral excess has involved the analysis of meteorites and in particular of the CC as just reported. In the present paper we address the problem of amino acid preservation in CC and the preservation of chirality. For the purpose of the present work we have selected 20 chiral α-amino acids, the common amino acids used in the biochemistry of the current terrestrial living organisms with the exclusion of glycine (the simplest but achiral α-amino acid). The 20 chiral amino acids selected were all in the l-enantiomeric form, since it is known that the current biochemistry of living organisms is using almost exclusively the l-enantiomers. The individual amino acids molecules were exposed to γ ray irradiation in closed vials to a total dose of 3.2 MGy (1MGy= 106 gray) in an approach similar to that used by us in the study of the radiation stability of C60 and C70 fullerenes (Cataldo, Strazzulla and Iglesias-Groth, 2009). Comets, asteroids and other bodies of the Solar System, especially those far away from the Sun the Kuiper belt or in the Oort cloud are thought to include very primitive organic material embedded in ices and rocks. Such material has not been processed so much since the times of the Solar System formation till the present and remained embedded inside the host bodies. The almost unique source of degradation and processing of the organic material occurring in 4.6x109 years (the Solar System age) at a depth >20 m in comets or asteroid derives almost exclusively from the radionuclide decay (Draganic et al. 1993). The Nobel laureate Harold C. Urey (1955, 1956) calculated the amount of energy generated by the decay of radionuclides in comets, asteroids, meteorites and larger bodies of the solar system. More recently, Draganic et al. (1993) reported results of such calculations essentially showing that in comets or asteroids at a depth of 20 meters or more, the cosmic rays are completely shielded and the produced radiation derives almost exclusively from the radionuclide decay. In a time scale of the age of the Solar System, i.e. 4.6x109 years (Unsold and Baschek, 2002), the total radiation produced by radionuclide decay in bulk comets, asteroids and larger bodies of the Solar System is ≈14 MGy. The short lived radionuclide 26Al is able to produce ≈11 MGy in the first billion years of existence of the Solar System (Kohman, 1997) while the remaining contribution to the total radiation dose of 14 MGy derives from the decay of radionuclide having long and very long decay time like for instance 40K,232Th, 235U, 238U (Urey, 1955 and 1956; Draganic et al. 1993). The calculations of Urey (1955, 1956) and Draganic et al. (1993) are in agreement also with other calculations made by Prialnik et al. (1987). In the present work we report about the decomposition and radioracemization of a series of selected amino acids which have been irradiated in the solid state at 3.2 MGy, a dose equivalent to that received by the aminoacids in comets or asteroids in 1.05x109 years. From the degree of decomposition and radioracemization reached we were able to establish both the relative resistance to radiation of each amino acid under study and to extrapolate the initial amount at the epoch of the Solar System formation. 2. Experimental 2.1 Materials and equipment A set of 20 proteinaceous amino acids shown in Scheme 1 (l-alanine [Ala], l-arginine [Arg], l- asparagine [Asn], l-aspartic acid [Asp], l-cysteine [Cys], l-cystine, l-glutamic acid [Gln], l- glutamine [Glu], l-histidine [His], l-isoleucine [Ile], l-leucine [Leu], l-lysine [Lys], l-methionine [Met], l-phenyalanine [Phe], l-proline [Pro], l-serine [Ser], l-threonine [Thr], l-tryptophan [Trp], l-tyrosine [Tyr], l-valine [Val]) were obtained from Sigma and used as received. The measurement of the degree of decomposition of the amino acids under study was made by two techniques: the Differential Scanning Calorimetric (DSC) analysis which was made on a DSC-1 Star System from Mettler-Toledo and the Optical Rotatory Dispersion spectroscopy performed on a Jasco P-2000 spectropolarimeter with a dedicated monochromator. 2.2 Irradiation procedure with γ rays The irradiation of the 20 amino acids with γ rays was made in a Gammacel from Atomic Energy of Canada at a dose rate of 1.7 kGy/h. A total dose of 3200 kGy = 3.2 MGy (1Gy = 1 J/kg) was delivered to each sample. The samples of each individual amino acids (300 mg each) were irradiated in the solid state insight tightly closed glass vials. 2.3 Analysis with differential scanning calorimetry The irradiated samples were tested for purity by a DSC (Differential Scanning Calorimetry) at a heating rate of 10°C/min in static air. As reference the DSC test was applied also on pristine, not irradiated samples under the same conditions. The amount % of residual sample after the solid state radiolysis Nγ was determined from the ratio of the melting enthalpy after the radiolysis at 3.2 MGy (∆Hγ) and the enthalpy before radiolysis measured on the pristine sample (∆H0): Nγ = 100 [∆Hγ / ∆H0] (1) 2.4 Analysis of the radioracemization degree by Optical Rotatory Dispersion spectroscopy. The degree of radioracemization was measured by ORD (Optical Rotatory Dispersion) spectroscopy. The irradiated amino acid samples were dissolved in HCl 1M and the optical rotation was measured on the resulting solution using a polarimetric cell of 0.5 dm length in the range between 400-600 nm. As reference, the same measurement was also made on standard pristine amino acid dissolved in the same medium at the same concentration. From the ratio of the specific optical rotation after radiolysis [α]γ and before radiolysis [α]0 the residual optical activity Rγ has been determined: Rγ = 100 [α ]γ / [α]0 (eq. 2) The value of Rγ was always averaged in the entire range of wavelengths (400-600 nm) where the specific optical rotation was measured. 3. Results and discussion 3.1 – The amino acids stability to radiation and radioracemization: the bottom-up approach The abiotic synthesis of amino acids starting from simple molecules like methane, ammonia, water occurs in different conditions when there is a sufficient source of energy from electrical discharges (Miller, 1953; 1955; 2000) to UV irradiation (Sagan and Khare 1971). The vacuum UV irradiation of interstellar ice analogs produces a mixture of amino acids and can be considered an extension of the Sagan synthesis (Bernstein et al. 2002; Nuevo et al. 2008). The irradiation with circularly polarized light of interstellar ices analogs produces the asymmetric synthesis of a mixture of amino acids (Muñoz Caro et al. 2002; Takano et al. 2007). Since there are known sources of circularly polarized light in the star forming regions (Bailey et al. 1998), it is thought that molecules like amino acids which are formed in racemic mixtures in prebiotic processes are transformed into scalemic mixtures (mixtures containing an enantiomeric excess) under the action of certain natural circularly polarized light sources. Then, these molecules are incorporated in comets and asteroids and other bodies of the solar system. Notably, as already mentioned, a series of amino acids some of them in chiral excess have been found in different meteorites (Pizzarello and Cronin, 2000; Sephton, 2002; Pizzarello et al. 2008). If adequately shielded from cosmic rays, the organic molecules present in asteroids and meteorites undergo radiolytic decomposition due to the decay of the radionuclides present in the rocks of those bodies. Thus, the questions we intend to answer are: what was the amount of amino acids present 4.6x109 years ago in the bodies just formed at the beginning of the Solar System? By determining the radiolysis half life period we will show that it is possible to answer to that question. Similarly can the enantiomeric excess present in the organic matter embedded in comets or asteroids (adequately shielded from cosmic rays) survive a continuous high energy irradiation for 4.6x109 years? It is well known that the high energy radiation plays always against the preservation of chirality (Cataldo, 2007b; Cataldo et al., 2008) because of the well known phenomenon of radioracemization. Only in exceptional cases the radiolysis of certain substrates may surprisingly lead to an enhancement of the chirality rather than to the radioracemization (Cataldo et al. 2007; 2009). Accepting the Urey (1955; 1956) and Draganic et al. (1991) calculation that the total radiation dose delivered in 4.6x109 years to the amino acids buried at a depth of >20 m in comets, asteroids and other bodies of the Solar System is estimated to be 14 MGy. From this, it can be deduced that the radiation dose used in our study: 3.2 MGy, corresponds to the dose equivalent to 1.05x109 years. It must be emphasized a limitation between of our experimental approach in comparison to real irradiation conditions derived from the radionuclide decay. In our case a dose rate of 1.7kGy/h was used so that we were able to reach the total dose of 3.2 MGy in 2.6 months. In this relatively short interval of time the dose rate remains constant. On the other hand in comets, asteroids and other bodies of the Solar System, the dose rate derived from the radionuclide decay is much lower and additionally it is not linear with time. For instance, it is well known that 26Al is completely consumed in less than 107 years at the very beginning of the story of the Solar System. In that relatively short time 26Al decaying by either of the modes beta- plus or electron capture, both resulting in the stable nuclide 26Mg, produces about 11 MGy of radiation dose (Draganic et al. 1993). All the other radionuclides (40K,232Th, 235U, 238U, 10Be, 129I, 237Np, 244Pu, 247Cm) produce about 3 MGy in 4.6x109 years. Thus, in the real cometary conditions, the dose rate was not linear at all, it was high at the beginning of the Solar System especially when there was the contribution derived from 26Al and was slowed down progressively by reaching the present time. In spite of these important differences in the dose rate between the experiment and the real conditions, in radiation chemistry the radiolysis effects are essentially governed by the total radiation dose (Hughes, 1973; Woods and Pikaev, 1994) and in this case the dose we have administered is 3.2 MGy. Of course we have to assume a linear relationship between the radiation dose and time. It is an artificial assumption which simplify our calculations but the final effects on the organic substrate are the same at the end. Thus, if the total radiation dose delivered by the radionuclides decay in 4.6x109 years is 14 MGy, then 3.2 MGy correspond to the dose delivered in 1.05x109 years. Once it has been determined the degree of decomposition of a given amino acid at 3.2 MGy, it can be extrapolated to 14 MGy, hence at 4.6x109 years. Assuming a first order kinetics in the solid state radiolysis of the amino acids we may calculate both the rate constant k and the half-life period T1/2 according to the following equations (Yeremin, 1979): Nγ = N0 e-kt (3) ln (N0/ Nγ) = kdsct (4) kdsc = [ln (N0/ Nγ)] t-1 (5) T1/2dsc = (ln2)/k = 0.963 k-1 (6) Where Nγ is the amino acid concentration at any time of radiolysis, N0 the concentration at the beginning of the radiolysis. Since N0 = 100 i.e. a given amino acid it is assumed 100% pure and Nγ the residual amount of a given amino acid as measured by DSC at 3.2 MGy, i.e. at 1.05x109 years, and as determined by eq. 1, it is possible to derive the value of k, the radiolysis rate constant from eq. 5. Once k is known, then it is possible to calculate the radiolysis half life of the amino acid. The amino acids under study were subjected also to Optical Rotatory Dispersion (ORD) spectroscopy with the purpose to study the radioracemization degree. To avoid artificially induced racemization, the pristine and solid state irradiated amino acids were dissolved in acidic or neutral solution (Djerassi, 1960; Jirgensons, 1973) before the ORD measurement. Thus, the pristine and the irradiated amino acid samples were dissolved in water or in HCl 1M at the same concentration and the ORD curve was measured with a spectropolarimeter. One thing is the radiolysis of a given amino acid without caring about its optical activity and its enantiomeric excess and another thing is instead the evaluation of the residual optical activity after the radiolysis at 3.2 MGy. By using the residual optical activity Rγ derived from the ORD curve, and making the same mathematical treatment adopted for Nγ, it is possible to derive the analogous equations: -1 (9) Rγ = R0 e-kt (7) ln (R0/ Rγ) = kract (8) T1/2rac = (ln2)/krac = 0.963 krac from where both the radioracemization rate constant krac and the half life of optical activity can be derived T1/2rac. 3.2 - The amino acids stability to radiation and radioracemization: the results The solid state irradiation of the amino acids causes the formation of free radicals in the molecular crystal followed by the decomposition of the pristine amino acid for instance by a deamination or a decarboxylation reaction (Sagstuen et al. 2004). Thus, the continuous radiolysis causes a steady decomposition but occurring in the solid state the radiolysis products are not so mobile and cannot escape easily, especially large molecular fragments remain trapped in the crystalline structure. With the DSC technique the measurement of the purity of a given crystalline compound is made through its melting enthalpy (Brown, 2001). Initially the melting enthalpy of a pristine, reference amino acid is measured and the onset and peak of the melting point determined as shown for example in Fig. 1 in the case of the amino acid alanine. Then, the same amino acid but after 3.2 MGy of radiation dose, is studied by DSC. As shown in Fig. 1, there is a significant reduction in the melting enthalpy of the irradiated sample and a shift of the melting point onset and peak. More in detail, pure alanine (Fig. 1) has a melting enthalpy of -1387.4 J/g and after the administration of 3.2 MGy it is reduced to -938.5 J/g. The reduction of the melting enthalpy is due to the partial decomposition of the alanine into other products and hence its purity is reduced to Nγ = 100 (-938.5/-1387.4) = 67.6%. The decomposition products do not participate to the melting enthalpy because do not form a crystalline structure and only the residual amino acid still possesses a crystalline structure, this justifies a reduction of the melting enthalpy. Similarly, the radiolysis products that remained trapped in the crystalline structure of the amino acids cause a reduction of the melting point. Thus, in the case of alanine shown in Fig. 1, the onset and peak melting point of the pristine, sample occur at 291.0°C and at 293.5°C respectively but after the radiolysis at 3.2 MGy such peaks transitions appear shifted at lower temperatures 275.7°C and 282.4°C respectively. Of course there is a correlation between the reduction of the melting enthalpy and the shift toward lower temperature of the melting transition. The other approach we have used to measure the degree of decomposition of all the amino acids was to measure their optical activity through the ORD measurement. The optical activity of any enantiomer is expressed through the Biot law (Djerassi, 1960; Jirgensons, 1973): [α]λ = α l-1c-1 (eq. 10) where the specific optical rotation [α]λ is a constant at a given wavelength and concentration c of the given optical active molecule; α is the rotation degree of the plane of polarized light after the passage through a polarimetric tube having an optical path length l. The value of [α]λ changes only with the wavelength. Thus, the optical rotatory dispersion curve (ORD) is the value of [α]λ as function of the wavelength as shown in Fig. 2 and 3. Pristine, reference amino acid at a certain concentration in HCl 1M give an ORD curve or spectra like those shown in Fig. 2 in the case of l-alanine and l-leucine or in Fig. 3 in the case of l-tyrosine, l-phenylalanine and l-tryptophan. When the solid state irradiated amino acids (at 3.2 MGy) are dissolved in HCl 1M at the same concentration of the corresponding reference pristine amino acid, produce an ORD spectrum which appear shifted down in the vertical direction. Since l-alanine and l-leucine give a positive specific optical rotation, the corresponding irradiated samples show a down shift of the entire ORD curve toward the abscissa axis (see Fig. 2). On the other hand, since the pristine amino acids l-tyrosine, l-phenylalanine and l-tryptophan show a negative specific optical rotation, the corresponding irradiated samples at 3.2 MGy dissolved in HCl 1 M at the same concentration show an upper shift of the ORD curve again in the vertical direction (see Fig. 3). In both cases such shift is a consequence of the radioracemization of the sample. According to eq. 2, Rγ is the residual amount of an amino acid enantiomer after 3.2 MGy radiation dose. The value of Rγ was determined at all the wavelengths shown in Fig. 2 and 3 and averaged. The radiolysis of amino acids may cause a simple inversion of the chiral centre of the enantiomer irradiated. In such a case we are dealing with a true radioracemization which involves the isomerization of the l-enantiomer into the d-enantiomer. Such an isomerization will imply a reduction in the intensity of the specific optical rotation at any wavelength. Because of the high energies involved in the solid state γ radiolysis of the amino acids does not involve only the true radioracemization but also the so-called “false” or apparent radioracemization. The “false” or apparent radioracemization involves the radiolysis of the given enantiomer into other products with the loss of any chiral centre. In that case, the reduction of the [α]λ value and the consequent shift of the ORD curve is not due to the transformation of one enantiomer into the other but to the decomposition of the enantiomer. The residual optical activity Rγ will hence be due to the residual enantiomer that survived the radiolysis. There is also another possibility: the survival of the chiral centre in the radiolysis products. Such an event is rare but possible and, of course, will lead in any case to an alteration of the ORD curve in such a way that it will appear as an apparent racemization. Table 1 summarizes all the experimental results regarding the solid state radiolysis of all proteinaceous amino acids studied both in terms of residual amino acid amount which survived the 3.2 MGy radiolysis measured by ORD spectroscopy (Rγ %, eq. 2) and measured by DSC (Nγ %, eq. 1). More in detail, Table 1 shows the proteinaceous amino acid ordered according to their radioracemization resistance from the least resistant to the most resistant. Threonine is the amino acid showing the lowest resistance to radioracemization and tyrosine is the most resistant. However, the most astonishing thing is that after 3.2 MGy radiation dose, a dose equivalent to 1.05x109 years of irradiation due to radionuclide decay, at least 62.8% of the original amount of l-threonine optical activity is preserved and the results are even more encouraging for all the other amino acids. Already some authors have reported about the surprising preservation of chirality to considerable radiation doses (Kminek and Bada, 2006; Cataldo 2007a; Cataldo et al 2008; Izumi et al. 2008) but the present study systematically examines the stability of all amino acids toward high energy radiation and the result is that practically in all cases they can survive an enormous radiation dose if administered in the solid state. Indeed Table 1 shows also a reasonable correlation between the amount of amino acids survived 3.2 MGy as measured from ORD spectroscopy (Rγ %) in comparison to the amount of the amino acids survived the radiolysis as measured by DSC (Nγ %). As already discovered by Bonner and Lemmon (1978), as a general rule to a certain degree of the radiolytic decomposition of the amino acids does not correspond necessarily exactly the same degree of radioracemization. TABLE 1 - AMINO ACIDS ORDERED ACCORDING TO THE RADIORACEMIZATION RESISTANCE Nγ % after 1.05x109 y 66.6 80.8 63.5 56.6 67.6 54.6 92.2 70.2 68.4 72.3 83.5 70.5 94.7 84.5 98.1 95.5 83.4 79.6 96.7 92.1 l-threonine l-serine l-cysteine l-arginine l-alanine l-lysine l-tryptophan l-phenylalanine l-methionine l-leucine l-proline l-isoleucine l-glutamine l-glutamic acid l-valine l-aspartic acid l-histidine l-asparagine l-cystine l-tyrosine Rγ % after 1.05x109 y 62.8 70.7 71.4 75.3 75.8 77.0 79.1 82.5 82.6 82.6 86.8 91.4 91.6 92.7 94.0 94.2 96.3 97.6 97.6 98.9 For example Table 1 shows that in terms of Nγ the worse radiolysis resistance belongs to lysine which however is ranked 15th in the radioracemization list (left column of Table 1). On the other hand the best radiolytic stability in the solid state (in terms of Nγ) is offered by valine which also shows a considerable radioracemization resistance but is ranked only 6th in the radioracemization list. These differences between the amount of amino acid survived the radiolysis as measured by DSC and expressed as Nγ and the amount of the same amino acids survived the radiolysis and measured by ORD spectroscopy and expressed by Rγ can be explained by advocating the true and apparent radioracemization phenomena discussed earlier. In any case Fig. 4 provides a reasonable correlation between the Nγ and Rγ parameter so that the following equation has been derived: Nγ = 0.864 Rγ + 5.182 (eq. 11) Although the R2 value of 0.452 demonstrates a certain dispersion of the data. Knowing the Nγ and Rγ values of each proteinaceous amino acid measured at 3.2 MGy which corresponds to 1.05x109 years of irradiation, we can extrapolate back these data to 4.6x109 years ago, i.e. the beginning of the Solar System. Table 2 provides a summary of such an extrapolation. TABLE 2 - RESIDUAL AMINO ACIDS AFTER EXTRAPOLATION TO 4.6 GIGAYEARS Rγ % after 4.6x109 y 13.1 21.9 22.9 28.9 29.7 31.9 35.8 43.1 43.3 43.3 53.8 67.5 68.1 71.8 76.3 77.0 84.8 89.9 89.9 95.3 Nγ % after 4.6x109 y 16.9 39.3 13.7 8.3 19.5 7.1 70.1 21.3 19.0 24.2 45.4 21.7 78.8 47.8 91.9 81.7 45.2 36.8 86.3 69.8 l-threonine l-serine l-cysteine l-arginine l-alanine l-lysine l-tryptophan l-phenylalanine l-methionine l-leucine l-proline l-isoleucine l-glutamine l-glutamic acid l-valine l-aspartic acid l-histidine l-asparagine l-cystine l-tyrosine Thus, in Table 2 the value of Rγ = 13.1% means that only such fraction of the original optical activity due to threonine survived from the beginning of the Solar System till the present time and, similarly only 16.9% of the threonine present at the beginning of the Solar System survived the radiolysis till the present day. However there are amino acids with considerable radiation resistance whose amount should be approximately the same as in the origin of the Solar System. It is the case of valine, or other amino acids like aspartic acid and glutamine which are expected to be today just 4/5 of the original amounts present at the beginning of the formation of the Solar System. The preservation of chirality is even more impressive: after 14 MGy administered in the solid state, the amino acids glutamic acid, valine, aspartic acid, histidine, asparagine, cystine and tyrosine still display >70% of the original chirality they had 4.6x109 years ago. Based on these data it is not at all a surprise that amino acids have been found in meteorites and in measureable chiral excess. Furthermore, this study supports experimentally the theory (Crovisier and Encrenaz, 2000) that amino acids were formed in the interstellar medium and in chiral excess and then have been incorporated in comets and asteroids at the epoch of the Solar System formation. Using eq. 5 and eq. 8 we have determined the radiolysis rate constant of each amino acid kdsc and the radioracemization rate constant krac in year-1 as reported in Table 3 respectively. As expected from the previous discussion on the Nγ and Rγ values, the kdsc and krac have the same order of magnitude ranging from 10-10 y-1 for the less radiation resistant amino acids to 10-11 y-1 for the most radiation resistant. Similarly also the half life of each amino acid has been determined using the eq. 6 and 9. Again the T1/2dsc and T1/2rac have the same order of magnitude in the range between 109 and 1010 year as reported in Table 3. The validity of our experimental approach can be checked y comparing the kdsc and T1/2dsc results derived by calorimetric measurements with those of Kminek and Bada (2006). These authors have irradiated alanine in the solid state with γ radiation and have measured the Nγ by HPLC analysis instead of the DSC used by us. For alanine they have found a radiolysis rate constant k = 3.43x10-10 y-1 in good agreement with the value found in the present work (Table 3) k = 3.56x 10-10 y-1. Kminek and Bada (2006) report also k = 5.26x10-10 y-1 for glutamic acid and k = 4.78x 10-10 for aspartic acid, while our results, as reported in Table 3 for these two amino acids, are respectively k = 1.60x 10-10 y-1 and k = 0.438x10-10 y-1, in fair agreement. TABLE 3 - RADIORACEMIZATION AND RADIOLYSIS RATE CONSTANTS AND RELATIVE HALF LIFES l-threonine l-serine l-cysteine l-arginine l-alanine l-lysine l-tryptophan l-phenylalanine l-methionine l-leucine l-proline l-isoleucine l-glutamine l-glutamic acid l-valine l-aspartic acid l-histidine l-asparagine l-cystine krac (x10-10 y-1) kdsc (x10-10 y-1) T1/2rac (x109 y) T1/2dsc (x109 y) 4.43 3.30 3.21 2.70 2.64 2.49 2.23 1.83 1.82 1.82 1.35 0.856 0.835 0.721 0.589 0.569 0.359 0.231 0.231 3.87 2.03 4.32 5.42 3.56 5.76 0.773 3.37 3.61 3.09 1.72 3.33 0.518 1.60 0.183 0.438 1.73 2.17 0.319 1.57 2.10 2.16 2.57 2.63 2.79 3.11 3.79 3.81 3.81 5.15 8.10 8.30 9.61 11.8 12.2 19.3 30.0 30.0 1.79 3.42 1.60 1.28 1.95 1.20 8.97 2.06 1.92 2.25 4.04 2.08 13.4 4.33 38.0 15.8 4.01 3.19 21.7 l-tyrosine 0.105 0.783 65.9 8.85 4. Conclusions The 20 proteinaceous amino acids studied can survive a radiation dose of 14 MGy, which is the dose estimated to be delivered along the life of the Solar System to all organic molecules present in comets or asteroids at a depth of 20 m or more. The dose of 14 MGy is due to the decay of all major radionuclides in 4.6x109 years. At a depth >20 m, the contribution from cosmic rays to this total dose is negligible because of the shielding. Our study shows that eachno acid has an individual response to high energy radiation but it is evident that all the proteinaceous amino acids can survive in relatively large amount to a radiation dose equivalent to that administered by the radionuclide decay from the beginning of the history of the Solar System to the present. Not only all the amino acids can survive to 14 MGy but also their chirality can be partially preserved even after such radiation dose. Therefore, based on these results it is not at all a surprise that amino acids have been found in meteorites and in measureable chiral excess. Furthermore, this study supports experimentally the theory (Crovisier and Encrenaz, 2000) that amino acids were present in the interstellar material that conformes the protostelar nebula and in chiral excess. Then they were incorporated in comets and asteroids at the epoch of the Solar System formation. With the experimental data available we can even predict the concentration of the amino acids in comets and in asteroids at the beginning of the Solar System and also we can extrapolate to the original chiral excess 4.6x109 years ago. Acknowledgements The present research work has been the Spanish Ministerio de Ciencia e Innovacion. References Bailey, J. Chrysostomou, A., Hough, J.H., Gledhill, T.M., McCall, A., Clark, S. Menard, F., Tamura, M., 1998, Science, 281, 672. Bernstein, M.P., Dworkin, J.P., Sandford, S.A., Cooper, G.W., Allamandola, L.J., 2002, Nature, 416, 401. Bonner, W.A., Lemmon, R.M., 1978, Bioorg. Chem., 7, 175. Brown, M.E., 2001, Introduction to Thermal Analysis: Techniques and Applications. Springer, Berlin, p. 215-227. supported by grant AYA2007-64748 of Cataldo, F., 2007a, Int. J. Astrobiol. 6, 1. Cataldo, F., 2007b, J. Radioanal. Nucl. Chem. 272, 82. Cataldo, F., Gobbino, M., Ursini, O., Angelini, G., 2007, J. Macromol. Sci. A Pure Appl. Chem., 44, 1225. Cataldo, F., Ursini, O., Angelini, G., 2008, Radiat. Phys. Chem., 77, 961. Cataldo, F., Ursini, O., Angelini, G., Ragni, P. 2009, J. Macromol. Sci. A Pure Appl. Chem. 46, 493. Cataldo, F., Iglesias-Groth, S., Manchado, A., 2009, Mon. Not. Roy. Astron. Soc. 394, 615. Cronin, J.R., Pizzarello, S., 1997, Science, 275, 951. Crovisier, J., Encrenaz, T., 2000, Comet Science. The Study of Remnants from the Birth of the Solar System. Cambridge University Press, Cambridge. Djerassi, C., 1960, Optical rotatory Dispersion Applications to Organic Chemistry. McGraw- Hill, New York. Chapter 15. Draganic, I.G., Draganic, Z.D., Adloff, J.P., 1993, Radiation and Radioactivity on the Earth and Beyond. CRC Press, Boca Raton, Florida. Engel, M.H., Macko, S.A., 1997, Nature 389, 265. Hughes, G., 1973, Radiation Chemistry. Clarendon Press, Oxford. Izumi, Y., Matsui, T., Koketsu, T., Nakagawa, K., 2008, Radiat. Phys. Chem. 77, 1160. Jirgensons, B., 1973, Optical Activity of Proteins and Other Macromolecules. Springer-Verlag, Berlin. Chapter IV. Jorissen, A., Cerf, C., 2002, Orig. Life Evol. Biosph. 32, 129. Kminek, G., Bada, J.L., 2006, Earth Planet. Sci. Lett. 245, 1. Kohman, T.P., 1997, J. Radioanal. Nucl. Chem. 219, 165. Kwok, S., 2009, Astrophys. Space Sci., 319, 5. Lough, W.J., Wainer, I.W. (eds), 2002, Chirality in Natural and Appled Science. CRC Press, Boca Raton, FL. Meierhenrich, U.J., 2008, Amino Acids and the Asymmetry of Life. Springer, Berlin. Meierhenrich, U.J., Nahon, L., Alcaraz, C., Bredehoft, J.H., Hoffmann, S.V., Barbier, B., Brack, A., 2005, Angew. Chem. Int. Ed. 44, 2. Meierhenrich, U.J., Munoz Caro, G.M., Bredehoft, J.H., Jessberger, E.K., Thiemann, W.H.P., 2004, Proc. National Acad. Sci., 101, 9182. Miller, S.L., 1953, Science, 117, 528. Miller, S.L., 1955, J. Am. Chem. Soc. 77, 2351. Miller, S.L., 2000, The endogenous synthesis of organic compounds. Chapter 3 in The Molecular Origin of Life: Assembling the Pieces of a Puzzle, edited by Brack A. Cambridge University Press, Cambridge, UK. Muñoz Caro, G.M., Meierhenrich, U.J., Schutte, W.A., Barbier, B., Arcones Segovia, A., Rosenbauer, H., Thiemann, W.H.P., Brack, A., Greenberg, J.M., 2002, Nature 416, 403. Nuevo, M., Auger, G., Blanot, D., d’Hendecourt, L., 2008, Orig. Life Evol. Biosph., 38, 37. Pizzarello, S., Cronin, J.R., 2000, Geochim. Cosmochim. Acta, 64, 329. Pizzarello, S., Huang, Y., Alexandre, M.R., 2008, Proc. National Acad. Sci. 105, 3700. Prialnik, D., Bar-Nun, A., Podolak, M., 1987, Astrophys. J. 319, 993. Sagan, C., Khare, B.N., 1971. Science, 173, 417. Sagstuen, E., Sanderud, A., Hole, E.O., 2004, Radiat. Res. 162, 112. Schmitt-Kopplin, P., Gabelica, Z., Gougeon, R.D, Fekete, A., Kanawati, B., Harir, M., Gebefuegi, I., Eckel, G., Hertkorn, N., 2010, Proc. National Acad. Sci., 107, 2763. Sephton, M.A., 2002, Nat. Prod. Rep., 19, 292. Takano, Y., Takahashi, J., Kaneko, T., Marumo, K. and Kobayashi, K., 2007, Earth Planet. Sci. Lett. 254, 106. Unsold, A., Baschek, B., 2002, The New Cosmos. An Introduction to Astronomy and Astrophysics. Springer, Berlin. Urey, H.C., 1955, Proc. National Acad. Sci., 41, 127. Urey, H.C., 1956, Proc. National Acad. Sci., 42, 889. Woods, R.J., Pikaev, A.K., 1994, Applied Radiation Chemistry. Wiley-Interscience, New York. Yeremin, E.N., 1979, The Foundation of Chemical Kinetics. Mir Publishers, Moscow. CH3 O NH2 OH NH N H NH2 [Ala] NH2 [Arg] O O O OH O OH NH2 NH2 O OH OH NH2 O NH2 SH OH O O O OH OH NH2 [Asn] O OH NH2 N N H [Asp] O OH NH2 NH2 [Cys] CH3 O NH2 [Ile] CH3 [Gln] [Glu] [His] NH2 O OH CH3 OH CH3 NH2 [Leu] [Lys] O NH2 OH S CH3 O OH NH2 O OH NH2 [Phe] O NH [Pro] OH OH O NH2 [Ser] [Met] OH O OH CH3 OH NH2 [Thr] O OH NH2 OH N [Trp] NH2 [Tyr] O OH CH3 O CH3 OH NH2 [Val] SCHEME 1 187-36 Alanine std, 6,4000 mg Integral -8879,27 mJ normalized -1387,39 Jg^-1 290,88 °C Onset Peak 293,48 °C 10 187-36 Alanine 3.2 MGy, 6,9000 mg W/g Integral -6475,49 mJ normalized -938,48 Jg^-1 275,66 °C Onset Peak 282,38 °C 50 100 150 200 250 300 °C Fig. 1 – DSC of alanine in Al crucible heating rate 10°C/min. The upper trace is due to standard pristine alanine and the lower trace is due to the 3.2 MGy treated sample. l-leucine std l-leucine 3,2MGy Alanine std Alanine 3,2 MGy 55 50 45 40 35 30 25 20 15 10 5 0 -5 N O I T A T O R L A C I T P O C I F I C E P S 375 400 425 450 475 500 525 550 575 600 625 WAVENUMBERS (nm) Fig. 2 – Optical rotatory dispersion (ORD) of leucine and alanine before and after the solid state radiolysis at 3.2 MGy. The extent of the radioracemization can be appreciated by the shift of the ORD curve toward the abscissa after the radiolysis. Concentration in HCl 1M (pristine and irradiated): alanine 4.1 mg/ml; leucine 5.2 mg/ml. N O I T A T O R L A C I T P O C I F I C E P S 0 -5 -10 -15 -20 -25 -30 -35 -40 -45 -50 -55 -60 -65 -70 Phenylalanine std Phenylalanine 3.2 MGy Tyrosine std Tyrosine 3,2 MGy Tryptophan std Tryptophan 3,2 MGy 375 400 425 450 475 500 525 550 575 600 625 WAVELENGTH (nm) Fig. 3 – Optical rotatory dispersion (ORD) of phenylalanine, tyrosine and tryptophan before and after the solid state radiolysis at 3.2 MGy. The extent of the radioracemization can be appreciated by the shift of the ORD curve toward the zero axis after the radiolysis. The radioracemization is minimal in the case of tyrosine and appreciable. Concentration in HCl 1M (pristine and irradiated): phenylalanine 5.0 mg/ml; tyrosine 10.0 mg/ml; tryptophan 8.0 mg/ml. N = 0,8636 R + 5,1822 R2 = 0,4516 r a e y a g i G 1 r e t f a % N 100 90 80 70 60 50 40 30 20 10 0 0 10 20 30 40 50 60 70 80 90 100 R % after 1 Gygayear Fig. 4 – Correlation between the residual amount of amino acid survived the radiolysis equivalent to 1 Gigayear as measured by DSC (N %) and amount measured by ORD (R %). There is correlation but the correlation index R2 is not particularly high, indicating a certain dispersion of the two set of data.
1105.4475
1
1105
2011-05-23T12:00:07
Collective dynamics of active cytoskeletal networks
[ "physics.bio-ph", "q-bio.SC" ]
Self organization mechanisms are essential for the cytoskeleton to adapt to the requirements of living cells. They rely on the intricate interplay of cytoskeletal filaments, crosslinking proteins and molecular motors. Here we present an in vitro minimal model system consisting of actin filaments, fascin and myosin-II filaments exhibiting pulsative collective long range dynamics. The reorganizations in the highly dynamic steady state of the active gel are characterized by alternating periods of runs and stalls resulting in a superdiffusive dynamics of the network's constituents. They are dominated by the complex competition of crosslinking molecules and motor filaments in the network: Collective dynamics are only observed if the relative strength of the binding of myosin-II filaments to the actin network allows exerting high enough forces to unbind actin/fascin crosslinks. The feedback between structure formation and dynamics can be resolved by combining these experiments with phenomenological simulations based on simple interaction rules.
physics.bio-ph
physics
Collective dynamics of active cytoskeletal networks Simone Kohler, Volker Schaller and Andreas R. Bausch Technische Universitat Munchen, Garching, Germany Self organization mechanisms are essential for the cytoskeleton to adapt to the requirements of living cells. They rely on the intricate in- terplay of cytoskeletal filaments, crosslinking proteins and molecular motors. Here we present an in vitro minimal model system consisting of actin filaments, fascin and myosin-II filaments exhibiting pulsative collective long range dynamics. The reorganizations in the highly dy- namic steady state of the active gel are characterized by alternating periods of runs and stalls resulting in a superdiffusive dynamics of the network's constituents. They are dominated by the complex compe- tition of crosslinking molecules and motor filaments in the network: Collective dynamics are only observed if the relative strength of the binding of myosin-II filaments to the actin network allows exerting high enough forces to unbind actin/fascin crosslinks. The feedback between structure formation and dynamics can be resolved by com- bining these experiments with phenomenological simulations based on simple interaction rules. Introduction The cytoskeleton of eukaryotic cells is a highly flexible and adaptable scaffold that undergoes constant remodeling to meet their changing needs. With its unique static and dynamic properties it facilitates tasks as com- plex and diverse as cell division, cell locomotion or phagocytosis. Intracellular patterns can emerge, as has been observed in the apical constriction during Drosophila gastrulation, where the tissue rearrangement is driven by pulsed contractions of the actin myosin cytoskeleton [1, 2]. Similar self organization mechanisms are of outmost importance in many aspects of cellular development [3]. All these processes rely on the intricate interplay between three major com- ponents: actin filaments, molecular motors and crosslinking proteins. While a polymer network consisting of filaments and crosslinkers result in a viscoelastic physical gel, molecular motors exert local forces and turn it into an active gel [4]. The dynamics in these active actomyosin gels can be coordinated in time and space as has been observed for the pulsed constrictions during dorsal closure giving rise to a collective behavior in vivo [5]. A processive force exertion of the molecular motors inside a purely viscous 1 environment would lead to ballistic motion of the cytoskeletal material, only limited by the viscous friction of the environment. Inside of a homogenous viscoelastic medium additional relaxation processes occur, which would mainly result in a slowdown of the dynamics. Yet, the cytoskeleton is far from being homogeneous. Here, the coordinated action of molecular motors and crosslinking proteins leads to the formation and stabilization of a wide variety of structures that in turn affect the large scale dynamics of the system. It is the feedback between structure formation and dynamic properties that sets the functional properties of cytoskeletal systems and active gels in general. However, due to the lack of adequate model systems, it remains difficult to address these feedback mechanisms. Force dissipation in the viscoelastic environment and the inherent hetero- geneity should modulate the dynamics in a nontrivial way: The transport by molecular motors relies on the interaction with discrete and anisotropic fila- ments, which could suggest the occurrence of anomalous diffusion processes. At the same time, the force exertion inside of crosslinked cytoskeletal networks are expected to induce a force rate dependent rupturing of the crosslinking points [6] and thus the locally induced changes of the elastic environment will also af- fect and modulate the transport and structure formation dynamics. Thereby, it would be conceivable that the connectivity of the filamentous network enables the coordination of the transport and reorganization processes resulting in the appearance of collective modes. To shed light on the principles underlying the physics of such active gels, to examine their microscopic dynamics and to classify the thereby resulting dy- namic structures, we study a reconstituted actin network that is actively set under stress by molecular motors [7]. While crosslinking proteins are required for the mechanical stability of the actin network, molecular motors introduce an active component to the networks [8, 9, 10, 11]. The presence of only motor filaments and ATP does not suffice to induce any reorganization or structure formation in an actin solution. Only the combination of motors and crosslinking molecules can result in structure formation [12] and even can induce a macro- scopic contraction of cytoskeletal in vitro networks [13, 14, 15]. Here we show that in an active and crosslinked network pulsative collec- tive modes develop, which depend critically on the interaction strength of the myosin-II motor filaments. Our recently introduced minimal approach [7] using a reconstituted actin/fascin network in presence of myosin-II filaments enables a backtracking of the pulsative collective dynamic properties to the structure formation: The connectivity of the network is the basis for the coordinated re- organization of the cytoskeletal structure in the steady state. The dynamics are dominated by the complex competition of crosslinking molecules and motor fil- aments in the network. The key player is the relative strength of the binding of myosin-II filaments to the actin network: only a high binding affinity allows ex- erting high enough forces to unbind actin/fascin crosslinks resulting in network activity. The resulting dynamics can be described by an anomalous diffusion of the network's constituents. 2 Results Active gels composed of actin filaments, the crosslinker fascin and skeletal muscle myosin-II filaments in presence of ATP undergo drastic struc- tural rearrangements. The dynamics within the network is characterized by a succession of persistent runs and stalling events of individual actin/fascin/myosin- II structures (fig. 1A-F and supplemental movie S1). In general, the individual structures move independently. Yet, for short time intervals of about 30 s they can also syncronize their movements. These pulsative collective modes are char- acterized by a coordinated movement in time and space of a few or even up to nearly all visible structures of the network (fig. 1G). Structures which are up to hundreds of microns apart are synchronized in their movement: during run phases their directional motion is coordinated (fig. 1C) and the active move- ment starts and stops almost simultaneously. This coordinated movement is not altering the steady state of the system and is not linked to a macroscopic contraction, which only occurs at much higher actin concentrations [7]. The co- ordinated reorganization is only possible if the structures are connected, which is a prerequisite for the observed long range nature of the interaction. Molecular motors can only transport and exert forces between filaments. Thus, it is the connectivity and discrete nature of the network together with the force exer- tion mechanism, which enables the collective motion to occur. Introducing the velocity cross-correlation function Iv(r) collective movements can directly be quantified. For each time point the average of the squared correlation function hIv(r)2ir reflects the level of correlation in the system. High values of hIv(r)2ir relate to a highly collective movement, while low values close to zero indicate intervals lacking collective modes (fig. 1G and fig. S1). At 0.1 mM ATP, for a correlation cutoff k = 0.15 about 9 % of the frames show correlated movement (see material and methods and movie S1). Microscopically, stall phases are the direct consequence of the presence of the crosslinking molecules while the runs can be ascribed to the activity of the myosin-II filaments. Accordingly, the dynamics and their collective modes should be set by the motor activity. Thus, lowering the motor activity by de- creasing the ATP concentration should lead to a continuous decrease of the network dynamics. Yet, the variation of the ATP concentration results in a more complex behavior, as the competition of the motors with the crosslinking molecules comes into play. We observe four distinct regimes (fig. 2): (i) in the absence of ATP a passive actin/fascin/myosin network consisting of small, dif- fusing clusters is observed (fig. 2A and supplemental movie S2). Their movement is not correlated (fig. 2E). The passive actin/fascin/myosin network structure is distinct from an actin/fascin bundle network. This can be explained by a competitive binding of fascin and myosin-II to the actin filaments. Moreover, myosin-II filaments seem to act as nucleation seed of the actin polymerization. (ii) At low ATP concentrations (10 µM, fig. 2B) the low motor activity is limiting the number of runs and the stalls of the network structures are dominating the low dynamics. Consequently, collective modes barely occur in this crosslinker dominated regime. (iii) In the intermediate regime (up to 0.1 mM, fig. 2C) the highest correlation is observed. (iv) At high ATP concentrations (fig. 2D and supplemental movie S3) correlated motion is again rarely observed. Moreover, 3 at high ATP concentrations the networks become disrupted and huge clusters emerge. The maximal area of these clusters typically exceeds the area of a compact disc with radius of 80 µm whereas structures at lower ATP concentra- tions remain significantly smaller. The structural changes and the subsequent decrease in the occurrence of pulative collective modes can be attributed to the lower affinity of ATP-myosin to actin filaments, which limits the overall proces- sivity of the myosin-II filaments. Thus the forces myosin-II filaments can exert are lower than at intermediate concentrations of ATP [18, 19]. As a consequence, forced unbinding is less likely at higher ATP concentration. Accordingly, large aggregates with many crosslinkers bound cannot be disrupted while small struc- tures with only a few crosslinking points are still ruptured and rearranged. As a consequence, material is transported from small to large and highly crosslinked structures leading to the observed cluster formation. As the material is accu- mulated in these clusters the connectivity between them decreases resulting in the observed low correlation. The dynamics of the individual structures in the network is best described by the mean square displacement, which is computed for about 1500 individ- ual points per sample. For Brownian diffusion, the mean square displacement, hr2(τ )i ∝ τ α, is expected to increase with time τ with a power law exponent α = 1 while 0 ≤ α < 1 or 1 < α ≤ 2 are indicative of sub- or superdiffusion, respectively [16, 17]. In presence of ATP, a clear superdiffusive behavior of in- dividual clusters is found: the mean square displacements increase in time with α > 1 (fig. 3A) [7]. This can be traced back to the complex alternation of runs and stalls of the individual network structures. The distribution of the power- law exponents depends on the ATP concentration (fig. 3B): The passive state (i) is characterized by a subdiffusive behavior which can be attributed to the network hindering the diffusion. Already at low ATP concentrations (ii), the dynamics become superdiffusive. This superdiffusivity is even more pronounced at intermediate ATP concentrations (iii). However, at high ATP concentrations (iv) again a lower superdiffusivity is observed and the width of the distribution increases. This width is originated in the highly heterogenous network organi- zation which predominantly occurs in this regime. Similarly to the elimination of pulsative collective modes, the lower superdiffusivity can be attributed to longer stalls (supplemental fig. S2) due to a decrease in forced unbinding events resulting in less activity and a larger variability. The maximal velocity of individual structures in run phases of 0.6 µm/s is observed for intermediate ATP concentrations while higher or lower ATP concentrations result in a decrease of the velocity in accordance with the results obtained for the pulsative collective modes or superdiffusivity. However, in all regimes the velocities in run phases are an order of magnitude lower than the velocity of myosin-II filaments observed in gliding assays at similar ATP concentrations [20]. Thus, in all cases the speeds are not limited by the maximal speed of the motors, but by the presence of crosslinking bonds. Alternatively to the variation of ATP concentration, the crossbridge strength of the myosin-II filaments can be altered independently by the ionic strength [21]. An increase of the KCl concentration up to 200 mM does not affect 4 the myosin-II filament length [22]. We also do not observe any differences in actin/fascin network structure upon variation of the ionic strength. Indeed, ex- ceeding a critical KCl concentration of 60 mM prevents any dynamic reorganiza- tion within the active gel (fig. 4): the structure of an active actin/fascin/myosin network at more than 60 mM KCl cannot be distinguished from a passive actin/fascin network. This critical concentration has also been observed in a gliding assay, where salt concentrations higher than 60 mM do not support motility of actin filaments [23]. At this ionic strength the affinity of ATP-myosin does not allow for permanent binding of the myosin-II to the actin filaments throughout its chemomechanical cycle. Thus, the observed network reorganization critically depends on the strength of the myosin-actin bond: Only a maximal binding strength of the myosin- II filaments to actin at low ionic strength allows exerting high enough forces to disrupt the actin/fascin crosslinking points and to reorganize the network structure. The network reorganization and the observed superdiffusivity not only rely on the competition between active and passive crosslinks but also on the respective binding strengths and affinities. The good accessibility of all relevant system parameters within this recon- stituted approach enables us to test the identified principles in a simple phe- nomenological simulation that is based on probabilistic interaction rules. Fascin bundles are modeled as polar rigid rods in a quasi two dimensional geometry. The bundles are propelled and crosslinked reflecting transport and passive bind- ing processes. Both active and passive binding processes are subjected to forced unbinding events [7]. The implementation of these basic processes already suffices to obtain a network dynamics strongly reminiscent to the experimental findings. Due to the concerted action of molecular motors, clusters are constantly transported and the trajectories they perform are characterized by a succession of runs and stalls (fig. 5A). During run phases the velocity of the fastest clusters does not exceed one tenth of the velocity of the motor proteins; and the larger the clus- ters are, the lower is their susceptibility to motor induced displacements and the less they move (fig. 5B). During stall phases clusters essentially exhibit a diffusive motion. In accordance with the experiment, runs and stalls average to a superdiffusive behavior, as can be seen in the distribution of exponents of the mean square displacement in fig. 5C. While this is in accordance with time dependent ensemble averages [7], the occurence of broad distributions directly relates to the intrinsic degree of heterogeneity in the system. Like in the exper- iment, the dynamics of individual clusters relies on local influencing factors like the local network connectivity, the local concentration of active crosslinks and the individual cluster size. The exact values of α depend on the details of the actin-fascin and actin- myosin interaction. In the simulation, this is governed by the motor and crosslinker on- and off-rates. An increase in the motor off-rate roff lowers the crossbridge strength and corresponds to the addition of KCl in the experiment. Similar to the experiment the increase of the motor off-rate hinders the reorga- nization in the system and individual structures, single rods or entire clusters 5 move less. This is reflected in a reduced exponent of the mean square displace- ment, which successively decreases with increasing motor off-rate (fig. 5D). If the motor activity drops below a critical value, the dynamics becomes subdif- fusive. In this regime the motor proteins are not strong enough to induce any significant network reorganization and the system essentially consists of one sin- gle cluster that only shows diffusive motion. A similar behavior is found if the on-rate of the crosslinking proteins pon is increased. The higher the on-rate, the more the cluster dynamics is shifted to stall phases and the exponent α decreases concomitantly (fig. 5D). Importantly, the simulation cannot retrieve the pulsative collective run and stall phases observed in the experiment, where remote structures move in a coordinated manner (fig. 1). This indicates that the experimentally observed dynamics not only depend on the identified microscopic interactions but are modulated by a weak but far reaching connectivity within the system. While the incorporation of active transport, crosslinking and forced unbinding suffices to explain the superdiffusive transport properties, the simulation cannot account for long range force percolation as the force balance is only local with a typical range of a rodlength. Discussion A minimal model system consisting of actin filaments, crosslink- ing molecules and myosin-II motor filaments exhibits drastic structural rear- rangements that in turn affect the dynamic properties of the system. This feedback between large scale structure formation and dynamics is set by the mi- croscopic details of the competition between myosin-II filaments and crosslinking molecules. The properties of the system are sensitively controlled by the bind- ing affinity of the myosin-II filaments to actin-fascin structures that directly regulates the maximal force the myosin-II filaments can exert. If the exerted forces are sufficiently high to disrupt actin/fascin structures, the system is dom- inated by the motor activity. In this regime, a highly dynamic network is formed. The dynamics in this state show pulsative collective modes which can be attributed to the connectivity of the polydisperse structures. The pulsative collective modes are not only found for directly connected structures but they are propagated throughout the network. They do not lead to a macroscopic contraction but persist in a dynamic steady state. The balance of motor to crosslinker strength alone sets not only the degree of dynamics in the network but also their coordination and range. A high motor activity and strength is necessary to maintain the connectivity in the network which ensures not only the occurence of pulsative collective modes but also a highly superdiffusive dynamics. Lowering the binding affinity of the motor heads limits the internal force exertion to the crosslinked actin filaments. Con- sequently, the crosslinking molecules dominate relative to the motor strength resulting in a non-collective dynamics with lower superdiffusivity and a crossover to normal diffusion in a highly heterogeneous network. This opens up the pos- sibility of inducing large scale reorganizations within an active gel, building local contractile elements or establishing the systems' heterogeneity by solely 6 changing the maximal force the motor proteins can exert. Any large scale reorganization within active gels depends on the intricate in- terplay between structure formation, aggregation and dynamic transport, which are regulated and determined by the here studied competition between force gen- erating and force bearing structures. For large scale reorganization of the intra- cellular cytoskeleton active transport of the cytoskeletal building blocks, their aggregation to higher order structure are essential for ensuing force generating capabilities. The dynamics of these processes will depend on the here stud- ied interaction and competition between the force generating and force bearing structures. The combination of simplified in vitro model systems with simula- tions based on simple interaction rules opens up the perspective of addressing the governing principles of the cytoskeletal structure formation processes. Materials and Methods Protein purification Myosin [24] and G-actin [25, 26] are extracted from rabbit skeletal muscle (rabbit meat was obtained from Hasenhof Weh, Moorenweis, Germany). To fluorescently label actin, actin is dialyzed against 50 mM boric acid (pH 8.0), 0.2 mM CaCl2, 0.2 mM ATP and polymerized by addition of 2 mM MgCl2 and 100 mM KCl. F-actin is incubated with Alexa Fluor 555 succinimidylester (Invitrogen) at a molar ratio of 1:1 at room temperature for 1 h and centrifuged at 100 000 g at 4 ◦C for 2 h. Labelled F-actin is depolymerized by dialysis against G-buffer (2 mM Tris/HCl (pH 8.0), 0.2 mM CaCl2, 0.2 mM ATP, 0.2 mM DTT) at 4 ◦C. This yields a 25 % degree of labelling. Recombinant human fascin is purified from E.coli BL21-CodonPlus- RP and stored at -80 ◦C in 2 mM Tris/HCl (pH 7.4), 150 mM KCl at 64 µM [27]. Fluorescence imaging Actin is polymerized at 1 µM by adding one-tenth of the sample volume of a 10-fold concentrated polymerization buffer (100 mM imidazole, 2 mM CaCl2, 30 mM MgCl2), 0.1 µM myosin-II, and 1 µM fascin, 20 mM creatine phosphate and 0.1 mg/mL creatine phospho kinase (Sigma) for ATP regeneration, 3 mg/mL casein, ATP and KCl at indicated concentrations. In this polymerization buffer, myosin readily polymerizes into filaments with a mean length of 0.6 µm independent of the KCl concentration. Samples are enclosed to hermetically sealed chambers to eliminate any drift in the network [7]. All data are acquired on a Zeiss Axiovert 200 inverted microscope with either a 10 x (NA 0.2) long distance objective or a 40 x (NA 1.3) oil immersion objective. Images are captured at 0.84 frames/s with a charge-coupled device camera (Orca ER, Hamamatsu) attached to the microscope via a 0.4 x camera mount. Image processing For tracing individual actin structures, images are background subtracted in ImageJ. To identify individual points in the network, an inten- sity threshold value is applied in ImageJ to generate a binary image. Individual structures are identified as more than 10 connected bright pixels. The structures are traced over time using the IDL tracking algorithm by John C. Crocker [29] for the intensity weighted centroid positions using Matlab R2008b (The Math- Works, Inc.). To minimize tracking artefacts, the trajectories are subjected 7 x(cid:3)/(cid:0)hv2ix − hvi2 to a gliding average over 4 frames. Subsequently, the trajectories are divided into "runs" and "stalls": Runs are identified as movements between two frames with velocities larger than 0.36 µm/s and a change in direction smaller than 45 ◦ and the lengths of continuous runs over 2 successive frames are calculated. Run velocities are calculated as mean velocity of a single run and mean val- ues are obtained by fitting Lorenz distributions. Mean square displacements are calculated for entire trajectories and the first 10 % of the resulting mean square displacements are used. The power exponent α is fitted logarithmically to mean square displacements longer than 5 time points. Average values of α are obtained by fitting a Gaussian to the distribution of the power exponent to account for the heterogeneity of the networks. To determine phases of correlated movement, the velocity cross-correlation func- tion of moving structures Iv(r) = (cid:2)hv(x+ r)v(x)ix − hvi2 x(cid:1) is eval- uated. The degree of correlation is proportional to the average of the squared correlation function hI 2 v (r)ir (fig. S1). The proportion of time points showing collective modes is evaluated by introducing a correlation cutoff k: Averaged squared correlation functions hI 2 v (r)ir above k are defined as collective modes; values below k are defined to relate to an uncorrelated movement. For the data displayed in fig. 1 k was set to 0.15. The variation of k only affects the absoulte values while the relative values and the trend are not affected. Simulations The experimental system consists of interconnected polar fascin bundles that are actively transported and set under stress by myosin-II fila- ments [7]. The microscopic processes that lead to the observed self organiza- tion, arise through the interplay of crosslinking events and active transport. More specifically, the competition between molecular motors and crosslinking proteins gives rise to the active movement, (re-)binding and forced unbinding. These microscopic interactions are the basis for numerical simulations. In a min- imal approach fascin bundles are modelled as monodisperse and polar rigid rods in a two dimensional geometry. If two filaments overlap active or passive bind- ing events occur based on probabilistic interaction rules governed by the motor and crosslniker on rates ron and pon. In a similar manner unbinding processes are calculated with the motor and crosslinker off-rates roff and poff . The dis- placements that arise due to the action of molecular motors are calculated based on generic velocity models similar to those presented in Ref. [28]. Both active and passive bindings are subjected to forced unbinding events: Conceptually the stress increase leads to rupture events within the cluster whereby especially weakly crosslinked structures with a low actin and/or crosslinker density are prone to rupture. In a simplified picture this can be modelled by increasing the off-rate with the number of motor proteins per individual cluster. If not indicated otherwise, all simulations were performed with a density ρ = 14/L2, a motor on-rate ron = 0.1, a motor off-rate roff = 9 · 10−4, a crosslinker on-rate pon = 9 · 10−4 and a crosslinker off-rate poff = 1 · 10−4. Acknowledgments We thank the DFG in the framework of the SFB 863, the German Excellence Initiative 'International Graduate School for Science 8 and Engineering' and 'Nanosystems Initiative Munich' for financial support. V.S. acknowledges support from the Elite Network of Bavaria by the graduate programme CompInt. References [1] Martin AC, Kaschube M, Wieschaus EF (2009) Pulsed contractions of an actin-myosin network drive apical constriction. Nature 457, 495 -- 9. [2] Martin AC, Gelbart M, Fernandez-Gonzalez R, Kaschube M, Wieschaus EF (2010) Integration of contractile forces during tissue invagination. J Cell Biol. [3] Karsenti E (2008) Self-organization in cell biology: a brief history. Nat Rev Mol Cell Biol 9, 255 -- 62. [4] Joanny JF, Juelicher F, Kruse K, Prost J (2007) Hydrodynamic theory for multi-component active polar gels. New J Phys 9, 422. [5] Solon J, Kaya-Copur A, Colombelli J, Brunner D (2009) Pulsed forces timed by a ratchet-like mechanism drive directed tissue movement during dorsal closure. Cell 137, 1331 -- 42 [6] Evans E, Ritchie K (1997) Dynamic strength of molecular adhesion bonds. Biophys J 72, 1541 -- 55. [7] Koehler S, Schaller V, Bausch AR (2011) Structure formation in active networks. Nat Mater doi:10.1038/nmat3009. [8] Humphrey D, Duggan C, Saha D, Smith D, Kas JA (2002) Active fluidiza- tion of polymer networks through molecular motors. Nature 416, 413 -- 6. [9] Smith D et al. (2007) Molecular motor-induced instabilities and cross link- ers determine biopolymer organization. Biophys J 93, 4445 -- 52. [10] Mizuno D, Tardin C, Schmidt CF, Mackintosh FC (2007) Nonequilibrium mechanics of active cytoskeletal networks. Science 315, 370 -- 3. [11] Koenderink GH et al. (2009) Liquids and structural glasses special feature: An active biopolymer network controlled by molecular motors. Proc Natl Acad Sci USA. [12] Backouche F, Haviv L, Groswasser D, Bernheim-Groswasser A (2006) Ac- tive gels: dynamics of patterning and self-organization. Phys Biol 3, 264 -- 73. [13] Kane RE (1983) Interconversion of structural and contractile actin gels by insertion of myosin during assembly. J Cell Biol 97, 1745 -- 52. 9 [14] Janson LW, Kolega J, Taylor DL (1991) Modulation of contraction by gelation/solation in a reconstituted motile model. J Cell Biol 114, 1005 -- 15. [15] Bendix PM et al. (2008) A quantitative analysis of contractility in active cytoskeletal protein networks. Biophys J 94, 3126 -- 36. [16] Metzler R, Klafter J (2000) The random walk's guide to anomalous diffu- sion: a fractional dynamics approach. Physics Reports, 1 -- 77. [17] Metzler R, Klafter J (2000) Accelerating brownian motion: A fractional dynamics approach to fast diffusion. Europhys Lett 51, 492 -- 98. [18] Cooke R, Bialek W (1979) Contraction of glycerinated muscle fibers as a function of the atp concentration. Biophys J 28, 241 -- 58. [19] Debold EP, Patlak JB, Warshaw DM (2005) Slip sliding away: load- dependence of velocity generated by skeletal muscle myosin molecules in the laser trap. Biophys J 89, L34 -- 6. [20] Kron SJ, Spudich JA (1986) Fluorescent actin filaments move on myosin fixed to a glass surface. Proc Natl Acad Sci USA 83, 6272 -- 6. [21] Brenner B, Schoenberg M, Chalovich JM, Greene LE, Eisenberg E (1982) Evidence for cross-bridge attachment in relaxed muscle at low ionic strength. Proc Natl Acad Sci USA 79, 7288 -- 91. [22] Katsura I, Noda H (1971) Studies on the formation and physical chemical properties of synthetic myosin filaments. Journal of Biochemistry 69, 219 -- 29. [23] Takiguchi K, Hayashi H, Kurimoto E, Higashi-Fujime S (1990) In vitro motility of skeletal muscle myosin and its proteolytic fragments. Journal of Biochemistry 107, 671 -- 9. [24] Margossian SS, Lowey S (1982) Preparation of myosin and its subfragments from rabbit skeletal muscle. Meth Enzymol 85 Pt B, 55 -- 71. [25] Spudich JA, Watt S (1971) The regulation of rabbit skeletal muscle contrac- tion. i. biochemical studies of the interaction of the tropomyosin-troponin complex with actin and the proteolytic fragments of myosin. J Biol Chem 246, 4866 -- 71. [26] MacLean-Fletcher S, Pollard TD (1980) Identification of a factor in con- ventional muscle actin preparations which inhibits actin filament self- association. Biochem Biophys Res Commun 96, 18 -- 27. [27] Vignjevic D et al. (2003) Formation of filopodia-like bundles in vitro from a dendritic network. J Cell Biol 160, 951 -- 62. [28] Liverpool D, Marchetti CM (2003) Instabilities of Isotropic Solutions of Active Polar Filaments. Phys. Rev. Lett. 90, 138102. 10 [29] Crocker, J. C., and Grier, D. G. Methods of digital video microscopy for colloidal studies. J Colloid Interface Sci 179, 298 -- 310 (1996). 11 Figure 1: Structure and dynamics of actin/fascin/myosin networks. Fluorescence micrographs (A -- E) of active actin networks at indicated times af- ter initiation of polymerization (0.1 mM ATP) show the dynamic reorganization within the network. Red arrows indicate the movement of individual points in the network with lengths 20-fold magnified and a time-average over three frames. Green arrows show the resulting overall movement in the field of view (lengths are 40-fold magnified). These long range reorganization processes are summa- rized in the colored time overlay (red to blue, F). Due to the low magnification used in this experiment, very small structures cannot be resolved. Thus, the connectivity between the structures is higher than can be seen in these fluores- cence micrographs. The trajectory of an individual structure exhibits persistent runs (inset, white line) intermitted by stall periods (inset, magenta dots). These run and stall phases are not only observed for individual structures, but also the whole network in the field of view exhibits pulsatile collective dynamics with movements being coordinated in time and direction: In stall phases, individual structures move predominantly for only short periods and in random directions (A,E). The stall phases are followed by periods with high activity in the entire network. During these run phases, the better part of the network shows long, persistent runs (B,C). The degree of coordination of the movement is measured by the average squared velocity cross-correlation function hI 2 v i, in which phases of collective movement are reflected as peaks. In such collective phases that last for up to 30 sec, the majority of the identified structures moves in approximately the same direction (G). 12 Figure 2: Dependence of the dynamics on the ATP concentration. In abscence of ATP, myosin-II filaments act as nucleation seeds, resulting in a multitude of small, well separated clusters as illustrated in the fluorescence micrograph 90 min after polymerization (A). In prescence of low (textbfB), in- termediate (textbfC) or high (textbfD) ATP concentrations, a highly dynamic network of larger structures is formed. The degree of correlated motion deter- mined by the percentage of frames correlated more than 0.15 is dependent on the ATP concentration: Only intermediate ATP concentrations allow for highly correlated movement (E). 13 Figure 3: Superdiffusivity in active actin networks. Mean square dis- placements of two individual structures at 2 mM ATP show different but su- perdiffusive power law exponents (A). The distributions of the mean square displacement powerlaw exponents α show maximal superdiffusivity at interme- diate ATP concentrations (B, 50 (cyan) to 100 µM (green); solid lines show gaussian fits). At low ATP concentrations (10 µM, blue) the low activity of myosin-II limits the superdiffusive behavior. At high ATP concentrations (0.5 (yellow), 1 (red) or 2 mM (magenta)), myosin-II filaments are not able to ex- ert high enough forces for constant network reorganization resulting in a lower mean superdiffusivity. Similarily, the highest mean run velocities are reached at intermediate ATP concentrations (C). Figure 4: Dependence of the dynamics on the KCl concentration. Above a critical KCl concentration of 60 mM (A), an active actin/fascin/myosin network cannot be distinguished from a passive actin fascin network (100 mM, B), while below the critical concentration an active network is formed (C). The fluorescence micrographs are taken 90 min after initiation of polymerization. The superdiffusivity as characterized by the mean mean square displacement power exponent α decreases above a critical KCl concentration of 60 mM (D). 14 Figure 5: Simulation results. (A) shows a simulation snapshot superim- posed with the time-resolved center-off mass positions of the respective clusters. As can be seen from the center-off-mass trajectories, individual clusters move in a succession of runs and stalls and in general move less the larger they are. This is reflected in the mean-square displacements depicted in (B): while the large cluster shows a diffusive motion (curve 3) the other clusters move in a superdiffusive manner with mean square displacement exponents α > 1. The heterogeneity in the system leads to a broad distribution of exponents α with a mean at α = 1.42 ± 0.24 (C). The exact value of the mean exponent depends on the details of the motor- and crosslinker (un-)binding kinetics. Increasing the motor off-rate roff leads to a decreased activity in the system, visible in the decline of the mean square displacement exponent α (D, blue curve). Likewise, the activity and hence the exponent α decline, if the crosslinker on-rate pon is increased (D, red curve). 15
1111.0069
1
1111
2011-10-31T22:35:21
Organization of the Bacterial Light-Harvesting Apparatus Rationalized by Exciton Transport Optimization
[ "physics.bio-ph", "quant-ph" ]
Photosynthesis, the process by which energy from sunlight drives cellular metabolism, relies on a unique organization of light-harvesting and reaction center complexes. Recently, the organization of light-harvesting LH2 complexes and dimeric reaction center-light harvesting I-PufX (RC-LH1-PufX) core complexes in membranes of purple non-sulfur bacteria was revealed by atomic force microscopy (AFM)1. Here, we report that the structure of LH2 and its organization within the membrane can be largely rationalized by a simple physical model that relies primarily on exciton transfer optimization. The process through which the light-harvesting complexes transfer excitation energy has been recognized to incorporate both coherent and incoherent processes mediated by the surrounding protein environment. Using the Haken-Strobl model, we show that the organization of the complexes in the membrane can be almost entirely explained by simple electrostatic considerations and that quantum effects act primarily to enforce robustness with respect to spatial disorder between complexes. The implications of such an arrangement are discussed in the context of biomimetic photosynthetic analogs capable of transferring energy efficiently across tens to hundreds of nanometers
physics.bio-ph
physics
  Organization of the Bacterial Light-Harvesting Apparatus Rationalized by Exciton Transport Optimization Elad Harel 1   Northwestern University, Department of Chemistry, 2145 Sheridan Road, Evanston, IL 60208 Correspondence and requests for materials should be addressed to Elad Harel 2145 Sheridan Road Technological Institute K344 Evanston, IL 60202 847-467-7580 [email protected]   Abstract 2   Photosynthesis, the process by which energy from sunlight drives cellular metabolism, relies on a unique organization of light-harvesting and reaction center complexes. Recently, the organization of light-harvesting LH2 complexes and dimeric reaction center-light harvesting I-PufX (RC-LH1-PufX) core complexes in membranes of purple non-sulfur bacteria was revealed by atomic force microscopy (AFM)1. Here, we report that the structure of LH2 and its organization within the membrane can be largely rationalized by a simple physical model that relies primarily on exciton transfer optimization. The process through which the light-harvesting complexes transfer excitation energy has been recognized to incorporate both coherent and incoherent processes mediated by the surrounding protein environment. Using the Haken-Strobl model, we show that the organization of the complexes in the membrane can be almost entirely explained by simple electrostatic considerations and that quantum effects act primarily to enforce robustness with respect to spatial disorder between complexes. The implications of such an arrangement are discussed in the context of biomimetic photosynthetic analogs capable of transferring energy efficiently across tens to hundreds of nanometers.   Introduction 3   The first step in photosynthesis is the absorption of light by the light-harvesting (LH) apparatus2. Transfer of energy from the LH to the reaction center (RC) leads to a stabilized charge-separated state across the membrane, which drives chemical transduction. The high symmetry of the pigment-protein complexes that compose the LH apparatus of bacterial photosynthesis as revealed by high-resolution x-ray crystallography3 has motivated extensive theoretical and experimental investigations in an attempt to understand both the biological4 and physical5 significance of the structure of these complexes and their organization within the membrane. The photosynthetic bacterium Rhodobacter sphaeroides serves as a model system to understand the functional role of this organization. The optical spectroscopy6-10 of individual pigment- protein complexes in these bacteria has revealed that both quantum and classical transport phenomena may play a role in the remarkable efficiency of energy transfer in these systems. Theoretical work on these complexes at a high level of detail9,11,12 has attempted to understand the influence of quantum effects on both intra- and intermolecular transfer within and between complexes. In this report, we examine the implications of assuming that the structure of these complexes is optimized for efficient photo-capture and transport. We use a simplified model that neglects quantum coherence and treats the coupling between pigments by simple dipole-dipole interactions only. By examining large structures of chromophores arranged in different topologies, we explain the most salient features of the LH apparatus: nonomeric symmetry of LH2, ring-like structure, inter-complex distance, inter-complex 4     topology, and inter-complex transfer time. Critically, our model takes into account only the most basic features of the complexes, allowing us to unambiguously identify the core principles responsible for its primary function – light harvesting and energy transfer. Furthermore, we demonstrate that quantum effects do not necessarily increase efficiency of transfer, but, rather, make transfer more robust to disorder – a critical feature of the LH apparatus necessary to operate efficiently in the ‘hot and wet’ environment of the organism. Finally, we briefly discuss approaches to mimic the high quantum efficiency and large exciton diffusion lengths that may significantly increase performance of current artificial photosynthetic constructs. Theory We consider a simple model in which an exciton system is coupled to a bath, ground state, and trap. This is essentially the Haken-Strobl model. We treat the problem by formalism nearly identical to the one used by Cho and Silbey in a recent publication13 in which they investigated simple linear chains and simple, nonlinear topological ! ! =   −[ℒ!"! + ℒ!"!!"# + ℒ!"#$% + ℒ!"#$ ]!(!) arrangements of loops with three and four sites. The master equation is given by where ! is the density matrix and ℒ are Liouville operators. The system consists of N (1) dipoles at positions, !! , including one trap state, ! = !", coupled by the dipole-dipole interaction, !!" . For the sake of simplicity the pigments will be assumed to lie in a plane and their dipoles will all lie parallel to one another, normal to the plane. While in LH2 this is not the case (dipoles are in a circular arrangement), the consequences for energy transfer and LH organization are negligible. Furthermore, while the membrane structure 5     of Rh. sphaeroides is a vesicle and not a plane, this effect is also negligible in the context of our analysis. A simpler arrangement of dipoles allows us to more easily transition to a discussion of possible designs for biomimetic systems that use repeating units of pigments. ignored since k! ≪   k ! for the systems examined here, where k ! represents the average For the remainder of this work, the decay part of the master equation will be trap rate and k! represents the relaxation rate to the ground state. ℒ!"#$ represents reaction center. It is expressed as [ℒ!"#$ ]!"   = !!" ,! + !!" ,! !! /2. The factor of two irreversible decay to a trapping state, e.g. a charge separated state in the photosynthetic comes from the fact that the theoretical limit for relaxation in a population is twice that of [ℒ!"##"$ ]!"   = !!"∗ . To simplify matters more, we make the stationary approximation in a coherence according to Redfield theory14. Finally, the dissipation is modeled as which the coherence terms are time-independent. More accurately, we assume that coherences dephase much more rapidly than the time scale of population relaxation, an approximation validated by recent experimental results on isolated LH2 complexes using coherent two-dimensional optical spectroscopy15. With this assumption, the master equation can be written as (Hnj ρjm − Hjmρnj ) − \x{FFFF}(1 − δnm )Γ∗nm + (kt δn,tr + kt δm,tr ) + i∆nm\x{FFFF} ρnm −i \x{FFFF}j \x{FFFF}=n,m 1 where !!" = !! − !! is the detuning and !! represents the energy of the ith site. H is the 2 (2) = ρnm δnm + iHnm (ρm − ρn ) system Hamiltonian in the site basis. This equation can be conveniently expressed in block matrix form as follows. 6     !! ! !!!! = !!0 0 ! where !! = (!! , !! , … , !!"!! , !!"!! , … , !! ) and !! ≡ !!!  describes the population at each (3) !! = (!!" , !!" , … , !!! , !!" , !!" , … , !!! , … , !!! , !!! , … , !!!!,! ) describes the coherence site, except for the trap state which is dealt with separately, and terms. Construction of the (N - 1) x (N2 - N) matrix, B, and the (N2 - N) x (N2 – N) matrix, !!! = !! and !! !! + !!! = 0. Solving these two equations gives K, are described in the appendix for sake of continuity. From (3), one gets that !! = −!! !!!! !! Integrating both sides and assuming that at very long times the population of the system !! 0 = !! !!!! ! has all decayed to the trap state, results in where ! = !! , !! , … , !!"!! , !!"!! , … , !! and !! ≡ !! (!) !! is the average residence by !!" = 1/!! (see Appendix). Solving for the average residence time at each site time of excitation at site n. At the trap state, the average residence time is simply given requires inversion of an (N 2- N) x (N2 – N) matrix, followed by inversion of an (N - 1) x (4) (5) (N - 1) matrix. Integrating out the time dynamics simplifies the analysis and computational demands considerably; For instance, for N = 100 sites the average residence time can be calculated in under 20 seconds using a quad-core workstation. The 1 !   ≈ 1 + !! ! quantum efficiency of transport to the trap is then given by !! where ! ≡ !!!! is the average trapping time. (6) 7     Classical transport may also be calculated using this basic formalism. For any N-site − !! ,!" !!" !! ,! !! − !! !! =   − ! system, the classical rate equations are given by !!! 2!!" !!"∗ + !!2 (!! ,!" + !! ,!" ) ! + !!" ! !!" ! !!" = where the classical rate constant is given by (8) In this context, !!" , is the line width of the transition from site i to site j. Again, one can solve this by simple matrix inversion to find the classical average residence time at each (7) site. Unlike the quantum case, which scales as the square of the system size, the classical rate equations scale linearly with the size of the system, therefore requiring an N x N matrix inversion. Results and Discussion The organization of this section is as follows. First, we show why ~9 fold symmetry is optimal for a ring-like structure of a given size (5 nm diameter in this case). Second, we show that nearly degenerate site energies lead to an optimal trapping time – strongly suggesting that utilizing identical chromophore units is advantageous. Next, we show that ring-like structures are advantageous in terms of structural disorder relative to linear chains or other topologies. This result is then related to the optimal organization of complexes in a staggered or packed, rather than linear arrangement. We also discuss the whether the organization of these complexes is highly tuned. Finally, we contrast purely classical and mixed quantum/classical transport.   8   The organization of the LH apparatus on spherical vesicles16 is shown in Figure 1. LH2 consists of two ring structures, named the B800 and B850 ring for their respective absorption bands in the near infrared (NIR). The diameter of a single, nonomeric LH2 complex is about 8 nm17. Ignoring the protein scaffold, the largest intra-ring distance for average trapping time as a function of the spectral detuning between sites, !!" = !! − either ring is about 5 – 5.5 nm. Using a genetic algorithm, we minimized the quantum !! , the dephasing matrix, ! ∗ , and the trap rate, !! , for three staggered, 5 nm rings containing N = 4 -14 elements each. The optimal energies and dephasing terms between coupled rings are shown in Figure 2. The coupling strength between site dipoles as given by the magnitude of the off-diagonal matrix elements is shown schematically by lines connecting sites. In each case, the donor and trap (i.e. acceptor) states were chosen so at to keep the transfer distance approximately constant. Besides the number of elements per ring, the topology remained the same for each case. A plot of the optimal average transfer time against the number of elements per ring shows that the minimum transfer time is reached around 8-9 elements per ring, with a tapering off at higher number of elements. From a biological perspective, utilizing the smallest number of elements that achieves the optimal performance minimizes the biochemical costs associated with creation of unproductive chromophores. It is also important to note that as the number of elements per ring reaches N > 6, each ring begins to act like a single entity. This stems from the fact that when the coupling between rings is stronger than the detuning between sites, the excitation becomes delocalized across the rings. The orientation of the dipoles within each complex, as long as it preserves circular symmetry, is therefore not terribly   important as this acts to modulate the intra-chromophore coupling but does not change 9   the dynamics or have noticeable influence on the organization of the LH apparatus. The delocalization feature is a critical element of the design strategy used by the LH apparatus as illustrated in the next section. We then explored the question of topology – whether a ring structure for LH2 was really compared to some other unexplored geometry. Consider the five rings shown in top panel of figure 3. Exciton diffusion through this large complex takes about 80.5 ps after optimization of the detuning, dephasing, and trap rate. The optimal trapping time is reached when the site energies are almost completely degenerate, while the optimal dephasing rate appears to vary from about 0 – 3 cm-1 (maps not shown). Now consider the same length structure, but this time with single sites instead of ring structures. Again, the optimal trapping rate is reached when the site energies are nearly degenerate and with very low values of the dephasing rate. This is in agreement with classical rate equations for linear chains, since the classical and quantum limits in this geometry are nearly identical18. The average transfer time for the linear chain is only 21.5 ps, about one quarter that of the ring structures. This suggests that linear chains are more efficient, which intuitively makes sense since the excitation spends less time on sites that do not directly participate in transfer. Now consider what happens when we incorporate spatial disorder, Δr, in the system. At 0.5 nm disorder (Δxmax = Δymax = 0.5 nm) in the plane, the transfer time in the linear chain increases by nearly five-fold to about 102 ps. For 1 nm disorder, it dramatically increases to 1.2 ns and with 2 nm disorder it reaches 6.2 ns. By stark contrast, in the ring chain, 0.5 nm disorder increases the transfer time by about 50%   to 113 ps, while 1 nm disorder increases it to only 136 ps. 2 nm disorder raises this 10   further only to 164 ps. We then repeated these runs multiple times (5 time for the ring and 20 times for the linear chains) to obtain better statistics and understand the general trends. Figure 4 shows the dependence of the degree of average trapping time on the spatial disorder. For the linear chain of sites, the dependence is almost perfectly matched with a cubic dependence on disorder. This may be explained by the cubic dependence of the dipolar interactions with distance between chromophores. The rate goes as the sixth power, but ‘diffusion’ in one- dimension goes as the square root of time, giving an effective cubic dependence. That is, each instance of random spatial disorder starting from an initial linear distribution of sites, gives an average distance between sites that scales in a manner similar to a quasi- one dimensional diffusing particle (i.e. sites can move in either direction, but initiate with a linear configuration). Also, notice that for the linear chain the error increases almost linearly with spatial disorder – another feature that may be explained by analogy to one- dimensional diffusion. In contrast, the dependence of the trapping time with spatial disorder is approximately linear in the regime examined for the ring case. These results clearly indicate that ring structures are more robust to disorder. Effectively, each ring acts like a single chromophore such that a given amount of disorder appears smaller to a large chromophore than to a smaller one. Therefore, it is advantageous to use large chromophores for the sake of robustness, at least up to a point. If a ring gets too large then the exciton diffusion length begins to suffer in proportion to the ring diameter. One may ask why not adopt some grid-like structure, or a ring that is partially filled in with   chromophores. As shown in Figure 5, such an arrangement wastes efficiency as 11   excitation is spent on sites that do not participate in direct transfer processes. Or to put it another way, crossing a ring scales as the diameter of the ring, while using the interior scales with the square. Therefore, a ring-like structure is optimal given the high symmetry needed to transfer energy in any given direction, combined with the inherent robustness utilizing larger effective units. The linear ring chain arrangement brings up another question of what is the best way to arrange the rings themselves with respect to one another. As can be appreciated from the analysis above, disorder inevitably creates regions in which the electrostatic coupling between two sites is too weak for efficient transfer, i.e. the rate limiting step comes about from “breaks” in the linear chain. Since each ring is itself acting like a single chromophore, the same argument can be made for a linear chain of rings as shown in Figure 6. However, if the rings are staggered, then disorder is less likely to produce such breaks. Or to put it more succinctly, a ring of rings is optimal for the same reason that the ring was optimal to begin with. However, a ring of rings does not provide a clear means by which to integrate a single reaction center, so a staggered or packed arrangement is a suitable compromise. Incorporating random disorder to the ring centers and less than 0.1 nm disorder to the elements of the rings themselves, demonstrates the robustness of the average trapping time in this organization (see Figure 3 and accompanying discussion). Disorder in the elements of the ring is small in large part because of the relatively stiff protein scaffold, which holds the chromophores in a precise arrangement due to electrostatic interactions with nearby amino acid residues. The   electrostatic interactions between the rings also act to increase the bandwidth of 12   absorption beyond that possible by individual ring. The energy spectra resulting from diagonalization of the system Hamiltonian for the packed and linear arrangements are also shown in figure 6 and qualitatively reproduce the energy spectrum observed in the B800 band of single LH2 complexes5,19. The approach we have used thus far is to optimize the minimal trapping time as a function of a relatively large parameter space involving the energies at each site and ~N2 dephasing terms between pairs of sites as well as the trapping rate. This brings up the question of whether natural photosynthetic systems have somehow tuned all of these parameters through an evolutionary process. Intuitively, it may be hard to believe that evolution could optimize so many parameters while still remaining functional. To explore this question in depth, we ran an optimization procedure for the 3-ring structure shown in figure 7. Repeated runs to insure proper convergence of our optimization scheme gave a value of 63 ± 2 ps average trapping time from donor at site 6 to acceptor to the same value and set all the dephasing terms to be equal: !!" = ! and !!" ∗ = !∗ . As at site 10. We then ran the same model (no optimization), but now fixed all the energies shown in Figure 7c, the optimal trapping time is reached at !∗ = !" cm-1 and !! = !!" cm-1, giving a trapping time of 71 ps, which is only 10% higher than the optimal value. This suggests that the system is not highly tuned and that only a two, rather than ~N2 parameters, are necessary to achieve near-optimal transport. These results indicates the a robust, artificial system may be fabricated without molecular-level control of the site   environment that would otherwise need tuning through specific interactions with a 13   tailored spectral density. Our results are also consistent with the results of other works20-25, in which finite dephasing assists the transport since it eliminates stationary states that do not couple to the trap state. It is also interesting to point out that the optimal trapping time configuration predicts about 4-5 picoseconds per site. Given the delocalization of the exciton, this number is consistent with photon echo measurements26 that show a ~5 ps decay of B850 band in intact membranes of Rh. Sphaeroides. Without adequate spatial or spectral resolution such spectroscopic measurements cannot straightforwardly distinguish decay at one site from decay of a delocalized state. These results demonstrate that this very simple model captures most of the salient features of the spectral and dynamical properties of the system without invoking quantum coherence, details of the high-level structure of the complex, or identification of the spectral density that couple the pigments to the bath. Next, we considered the role of quantum mechanics in the exciton energy transfer compared to purely classical transport mechanisms for the 3-ring structure shown in figure 8. Interestingly, the quantum model shows a slower optimal transfer time by about a factor of three when compared to the optimal solution found using the classical model (red and blue arrows in the inset plot). At first glance, it appears then that quantum mechanics does not aid in transfer. However, when considering the issue of robustness a more complicated picture emerges. Again, consider the case in which the site energies 14     are degenerate and the dephasing rate (i.e. line width for the classical case) between all sites is identical (Figure 8A). The optimal transfer time in the classical picture is at near- zero values of the line width and at a trap rate of ~125 cm-1. Again, this picture is consistent from consideration of the denominator of equation (8). In the quantum case, the optical transfer time is reached at about the same trap rate, but at a finite dephasing rate of about 10 cm-1. As before, tuning only these two parameters produces nearly the same transfer times as the full optimization. Now, consider the case when there is energetic disorder in the system (Figure 8B). Here, we have added a random detuning up to a maximum of 100 cm-1 at each site. In the quantum case, the trapping time map is remarkably similar in the presence of this type of disorder. However, for the classical case, the line width shifts from near-zero in the degenerate case to about 33 cm-1 in the disordered case. Changes in site energies cause dramatic changes in the optimal line width for purely classical transport, while for quantum transport the system is significantly more robust. From a biological perspective this strategy is advantageous because the dephasing rate / line width is set by the details of the protein environment. Quantum mechanics acts to overcome changes in the energy landscape of the system that appear as kinetic traps in classical transport phenomenon, primarily through delocalization and through mechanisms that allow transport even between sites that are not directly coupled. These non-local kinetic networks were recently explored by Cho and Silbey for simple, linear chains and three- and four-level non-linear topologies13. Of course, the LH2 complex is composed of two rings, not one. Therefore, it is worth commenting on the possible origins and benefits of incorporating two rings. Purple non-sulfur bacteria do not utilize a chlorosome antennae complex as do green   sulfur bacteria. The chlorosome consists of many tens to hundreds of thousands of 15   photosynthetic pigments, but is devoid of a protein matrix. The purpose of the chlorosome is simply to act as a broadband and large surface area antenna2 to capture solar flux and direct it to the reaction center through an intermediate pigment-protein complex called the Fenna-Mathews-Olson (FMO) complex. The chlorosome funnels energy to populate one specific state, delocalized about ~2 sites, in the FMO complex. In the absence of the chlorosome, the LH apparatus of purple bacteria utilizes localized and broadband antennae – namely, the B800 rings. These antennae may simply serve the same purpose as the chlorosome, but with a distributed rather than localized architecture. Energy absorbed by the B800 ring transfers to the B850 ring in about 1 ps and is then directed towards a single acceptor site. Due to the strong intra-chromophore coupling in the B850 ring combined with static and dynamic disorder in the complex energy rapidly (< 300 fs) localize on one specific site. From this point forward, transfer may be described by the model used in this work. The origin of whether or not to utilize a chlorosome to begin with, however, is most likely driven by other biological considerations that have little or nothing to do with energy transfer. Finally, we discuss the implications for energy transfer in biomimetic systems based on these results. The fact that fine-tuning is not a necessary feature for near- optimal transfer is promising; using repeating units with nearly degenerate site energies and nearly equal dephasing rates is significantly easier than controlling the system-bath interactions at each site. Depending on the synthetic method used to piece together individual chromosomes into large units such as rings, the relative orientation of the transition dipoles may vary. Creating rings with different properties may be achieved by   capitalizing on relative dipole orientations or by using of different constituent dipoles, 16   number of elements, and local scaffold environment. Fortunately, the exact details of the chromophores are less important than the delocalized states that they form. Therefore, many candidates may be utilized as chromophores and trap states (i.e. charge separated species). Controlling the system-bath interactions – the dephasing rates in our model – may be achieved by using a polymer host with varying degree of cross-linking and, therefore, stiffness. Such constructs are yet to be explored in this context. Conclusion We have shown that the most salient features of the organization of the LH apparatus in model bacterium can be explained with the Haken-Strobl model without consideration of the high-level details of LH2 and LH1-RC complex. Integrating out the dynamics and considering only the mean residence time at each site simplifies the analysis considerably, allowing us to focus on the basic physics. By utilizing an optimization scheme, we have shown that rings are ideal structures by which to transfer energy efficiently for several reasons: 1) the high symmetry allows transfer in any direction, which is critical when the location of the RC (i.e. trap) is not known a priori, 2) the ring essentially acts to make the chromophore larger, increasing robustness to disorder among complexes, 3) filling in the interior of the structure acts to waste precious biochemical resources by spending time at sites that do not participate in efficient transfer across space; hence, a void inside the structure is ideal. Taken together, these features uniquely define a ring. We have also shown that a composite of rings in a staggered or packed configuration is more robust to spatial disorder as it avoids the likelihood of breaks that   are the rate-limiting step in the transfer process. We demonstrate that while our 17   optimization schemes may give the lowest transfer time, the system need not be fine- tuned to reach a comparable transfer time. This observation makes designing artificial photosynthetic complexes much more promising than if fine-tuning was necessary to achieve high quantum efficiency. Nature’s strategy to utilize repeating units of identical chromophores and place them in a single protein host is advantageous from a biological as well as physical perspective. Finally, we showed that quantum mechanical effects, at least based on this simplified model in which quantum coherence or pathway interference is effectively ignored, does not necessarily aid in transfer efficiency. Rather, it increases the robustness of the system to a fixed value of the dephasing and trap rates. Natural systems cannot dynamically tune these parameters in response to changes in energetic disorder since they are inherent to the pigment-protein structure. From a biological point- of-view, robustness is oftentimes more important than overall efficiency, especially when the decay to the ground state is so much slower than the time scale of transfer. Remarkably, the crude model employed here, which neglects the high-order structure, predicts the correct decay times observed in nonlinear spectroscopic experiments between LH2 complexes. The energy spectrum also matches qualitatively with recent single- molecule spectra of LH2 at low temperature. In this work we demonstrate that general features of organization are more important than the fine details of the LH apparatus, which is very promising in terms of designing artificial photosynthetic analogs.   Appendix. !!" ,!" = −!!! ,! Construction of K: !!" ,!" = !!! ,! !!" ,!! = −!!"∗ + !!!" + 12 (!! !! ,!" + !! !! ,!" ) 12, 13, … , 1! , 21, 23, … , 2! , … , !1, !2, … , ! − 1 ! → 1,2, … , ! ! − !. where the indexes of K are constructed by the one-to-one mapping 18   (A1) (A2) (A3) Construction of B: ! ! = !!" ⋯ !!,!!!! ⋮ ⋱ ⋮ Let B’ be an (N - 1) x (N2 – N) matrix, defined as follows !!" ⋯ !!,!!!! (A4) ! = ! or ! = ! . In the former case, the sign of the nm element is ‘+’, while in the latter it B is constructed by retaining only the elements, nm, of the ith row of B’ in which either is ‘-‘. All other elements are set to zero. Expression for !!" : From (2), the expression for !!" contains an additional term !! !!" . Therefore, after time 1 = k !τ!" . integration and assuming that all the excitation eventually reaches the trap, we get   Figures 19     Figure 1: Light-harvesting apparatus in purple bacteria. Illustration of spherical chromatophore vesicle from R. sphaeroides showing organization of light harvesting complexes, LH2, and light-harvesting - reaction center complex (LH1-RC). Architecture and arrangement of constituent chromophores based on AFM images. Figure used with permission from Sener et al.16   20     Figure 2. Optimal number of elements per ring as found by a genetic algorithm. Average trapping time as a function of the number of elements, N, in each ring. The diameter of each ring is 5 nm. The top row of images shows the strength of electrostatic coupling (lines between sites). Dashed lines correspond to > 5 cm-1 and < 10 cm-1, dotted lines to > 10 cm-1 and < 20 cm-1, and thick lines to > 20 cm-1. Transition dipole at each site is normal to the plane. Color of connecting lines indicates the dephasing rate between two sites. Color of circles at each site corresponds to their energies. The bottom row of images shows the average residence time at each site, indicated by the color of the     circles. The average trapping time is the sum of the average residence times at each site. The donor and trap states are labeled.     21     Figure 3: Chain of rings versus chain of individual chromophores. Comparison of the disorder between rings (left) and between chromophores (right) in a linear arrangement.     Total distance from donor to trap is approximately the same in each case (~32 nm). ∆r is the maximum, random displacement in both the x and y direction.   22     Figure 4: Rings are robust to spatial disorder. Plot of average trapping time versus static disorder between rings (red) and single sites (blue) from figure 3. Error bars for ring arrays are based on five runs through the optimization code. Error bars for array of single sites based on twenty runs through the optimization code. Outliers larger than 5σ were removed from the analysis. Single site arrays were fit to third-order polynomial, while ring arrays were fit to a least squares regression line. Inset shows a narrower window of trapping times (maximum of 400 ps). Arrows on the left of the inset indicate the mean residence time with no spatial disorder between rings. The green bar indicates approximate excited-state lifetime of Bacteriochlrophyll a, which represents the upper limit of relaxation of the sites back to the ground state. The linear array of sites performs     better than the linear array of rings with no spatial disorder, but is significantly less robust to imperfections in positioning.   23     Figure 5: Grid-like array of chromophores is highly inefficient. Excitonic transfer through a grid-like arrangement of sites after optimization. Left: Colors of circles represent site energies. Color of lines represent dephasing rate. Right: Color represent mean residence time at each site. A significant amount of time is “wasted” at sites that do not directly link the donor to acceptor, i.e. sites 7, 13, and 19.   24   Figure 6: Two-dimensional packing is more robust to spatial disorder than linear chains. Electrostatic coupling between staggered (left) versus linear (right) in the presence of spatial disorder between rings for a 7-ring system. Thin lines indicate weak coupling (> 5 cm-1 and < 10 cm-1), medium lines indicate intermediate coupling (> 10cm-1 and < 20 cm- 1) and thick lines indicates strong coupling (> 20 cm-1). Staggered arrangement maintains non-negligible coupling strength and hence a path from donor to acceptor through multiple, neighboring rings. Linear arrangement more easily forms breaks, which may effectively block energy transfer across large distances. Bottom right: spectrum calculated by diagonalizing the system Hamiltonian. Energy spans approximately 200 cm-1 in each case – a major role of coupling is to break site degeneracy and broaden the spectrum for efficient absorption of solar flux.   25   Figure 7: Near-optimal tapping is achieved without fine-tuning the system and bath. Left: Exciton transfer optimization achieved by a genetic algorithm to minimize the average trapping time as a function of the mutual dephasing between sites, trapping rate, and site energies. In the case of three rings in this arrangement, the optimal trapping time was found to be 63 ps. Right: Keeping the dephasing between sites constant and the site energies identical, the optimal trapping time is found to be ~70 ps. This indicates that fine-tuning of the system and the system-bath interactions is not necessary to achieve near-optimal transfer efficiency.   26   Figure 8: Quantum transport is robust to static energetic disorder. Comparison of classical and quantum transport in the absence (A) and presence (B) of static energetic disorder – difference in energies at each site. A. When the site energies are degenerate, classical transport predicts a shorter trapping time – near zero line width. Quantum transport predicts a slower trapping time by about a factor of three, but at a modest value of the dephasing rate. B. When the site energies are non-degenerate, classical transport undergoes a dramatic shift in the optimal line width. For the quantum case, changes in the optimal dephasing rate and trapping rate are negligible. In this case, the quantum transport is faster and significantly more robust to changes in site energies.   References 1 27   Bahatyrova, S. et al. The native architecture of a photosynthetic membrane. Nature 430, 1058-1062, doi:DOI 10.1038/nature02823 (2004). Blankenship, R. E. Molecular mechanisms of photosynthesis. (Blackwell Science, 2002). MCDERMOTT, G. et al. CRYSTAL-STRUCTURE OF AN INTEGRAL MEMBRANE LIGHT-HARVESTING COMPLEX FROM PHOTOSYNTHETIC BACTERIA. Nature 374, 517-521 (1995). PFENNIG, N. PHOTOSYNTHETIC BACTERIA. Annual Review of Microbiology 21, 285-& (1967). Cogdell, R., Gall, A. & Kohler, J. The architecture and function of the light- harvesting apparatus of purple bacteria: from single molecules to in vivo membranes. Quarterly Reviews of Biophysics 39, 227-324, doi:DOI 10.1017/S0033583506004434 (2006). Ketelaars, M. et al. Spectroscopy on the B850 band of individual light-harvesting 2 complexes of Rhodopseudomonas acidophila I. Experiments and Monte Carlo simulations. Biophysical Journal 80, 1591-1603 (2001). Matsushita, M. et al. Spectroscopy on the B850 band of individual light- harvesting 2 complexes of Rhodopseudomonas acidophila II. Exciton states of an elliptically deformed ring aggregate. Biophysical Journal 80, 1604-1614 (2001). Herek, J. L. et al. B800-->B850 energy transfer mechanism in bacterial LH2 complexes investigated by B800 pigment exchange. Biophys J 78, 2590-2596 (2000). Mukai, K., Abe, S. & Sumi, H. Theory of rapid excitation-energy transfer from B800 to optically-forbidden exciton states of B850 in the antenna system LH2 of photosynthetic purple bacteria. Journal of Physical Chemistry B 103, 6096-6102 (1999). Hess, S. et al. Femtosecond Energy-Transfer within the Lh2 Peripheral Antenna of the Photosynthetic Purple Bacteria Rhodobacter-Sphaeroides and Rhodopseudomonas-Palustris Ll. Chemical Physics Letters 216, 247-257 (1993). Borisov, A. Y. Discrepancy between experimental and theoretical excitation transfer rates in LH2 bacteriochlorophyll-protein complexes of purple bacteria. European Biophysics Journal with Biophysics Letters 37, 143-151, doi:10.1007/s00249-007-0200-0 (2008). Krueger, B. P., Scholes, G. D. & Fleming, G. R. Calculation of couplings and energy-transfer pathways between the pigments of LH2 by the ab initio transition density cube method (vol 102B, pg 5384, 1998). Journal of Physical Chemistry B 102, 9603-9603 (1998). Cao, J. S. & Silbey, R. J. Optimization of Exciton Trapping in Energy Transfer Processes. Journal of Physical Chemistry A 113, 13825-13838, doi:Doi 10.1021/Jp9032589 (2009). REDFIELD, A. ON THE THEORY OF RELAXATION PROCESSES. Ibm Journal of Research and Development 1, 19-31 (1957). Harel, E. & Engel, G. S. (submitted, 2011). 2 3 4 5 6 7 8 9 10 11 12 13 14 15 19 20 21 18 17   16 28   Sener, M., Olsen, J., Hunter, C. & Schulten, K. Atomic-level structural and functional model of a bacterial photosynthetic membrane vesicle. Proceedings of the National Academy of Sciences of the United States of America 104, 15723- 15728, doi:DOI 10.1073/pnas.0706861104 (2007). Olsen, J. et al. The Organization of LH2 Complexes in Membranes from Rhodobacter sphaeroides. Journal of Biological Chemistry 283, 30772-30779, doi:DOI 10.1074/jbc.M804824200 (2008). Briggs, J. & Eisfeld, A. Equivalence of quantum and classical coherence in electronic energy transfer. Physical Review E 83, -, doi:ARTN 051911 DOI 10.1103/PhysRevE.83.051911 (2011). Hofmann, C., Aartsma, T. J. & Kohler, J. Energetic disorder and the B850-exciton states of individual light-harvesting 2 complexes from Rhodopseudomonas acidophila. Chemical Physics Letters 395, 373-378, doi:Doi 10.1016/J.Cplett.2004.08.020 (2004). Rebentrost, P., Mohseni, M. & Aspuru-Guzik, A. Role of Quantum Coherence and Environmental Fluctuations in Chromophoric Energy Transport. Journal of Physical Chemistry B 113, 9942-9947, doi:DOI 10.1021/jp901724d (2009). Rebentrost, P., Mohseni, M., Kassal, I., Lloyd, S. & Aspuru-Guzik, A. Environment-assisted quantum transport. New Journal of Physics 11, -, doi:ARTN 033003 DOI 10.1088/1367-2630/11/3/033003 (2009). 22 Mohseni, M., Rebentrost, P., Lloyd, S. & Aspuru-Guzik, A. Environment-assisted quantum walks in photosynthetic energy transfer. Journal of Chemical Physics 129, 174106, doi:Artn 174106 Doi 10.1063/1.3002335 (2008). Caruso, F., Chin, A., Datta, A., Huelga, S. & Plenio, M. Entanglement and entangling power of the dynamics in light-harvesting complexes. Physical Review a 81, -, doi:ARTN 062346 DOI 10.1103/PhysRevA.81.062346 (2010). Caruso, F., Chin, A. W., Datta, A., Huelga, S. F. & Plenio, M. B. Highly efficient energy excitation transfer in light-harvesting complexes: The fundamental role of noise-assisted transport. Journal of Chemical Physics 131, 105106, doi:Artn 105106 Doi 10.1063/1.3223548 (2009). Plenio, M. B. & Huelga, S. F. Dephasing-assisted transport: quantum networks and biomolecules. New Journal of Physics 10, 113019, doi:Artn 113019 Doi 10.1088/1367-2630/10/11/113019 (2008). Agarwal, R. et al. Nature of disorder and inter-complex energy transfer in LH2 at room temperature: A three pulse photon echo peak shift study. Journal of Physical Chemistry a 106, 7573-7578, doi:DOI 10.1021/jp014054m (2002). van Oijen, A., Ketelaars, M., Kohler, J., Aartsma, T. & Schmidt, J. Spectroscopy of single light-harvesting complexes from purple photosynthetic bacteria at 1.2 K. Journal of Physical Chemistry B 102, 9363-9366 (1998). 23 24 25 26   27
1305.5094
2
1305
2013-06-27T08:50:58
Muscle as a meta-material operating near a critical point
[ "physics.bio-ph" ]
Passive mechanical response of skeletal muscles at fast time scales is dominated by long range interactions inducing cooperative behavior without breaking the detailed balance. This leads to such unusual "material properties" as negative equilibrium stiffness and different behavior in force and displacement controlled loading conditions. Our fitting of experimental data suggests that "muscle material" is finely tuned to perform close to a critical point which explains large fluctuations observed in muscles close to the stall force.
physics.bio-ph
physics
APS/123-QED Muscle as a Metamaterial Operating Near a Critical Point M. Caruel,1, 2 J.-M. Allain,2 and L. Truskinovsky2, ∗ 1Inria, 1 rue Honor´e d'Estienne d'Orves, 91120 Palaiseau, France 2LMS, CNRS-UMR 7649, Ecole Polytechnique, 91128 Palaiseau Cedex, France (Dated: May 31, 2021) The passive mechanical response of skeletal muscles at fast time scales is dominated by long range interactions inducing cooperative behavior without breaking the detailed balance. This leads to such unusual "material properties" as negative equilibrium stiffness and different behavior in force and displacement controlled loading conditions. Our fitting of experimental data suggests that "muscle material" is finely tuned to perform close to a critical point which explains large fluctuations observed in muscles close to the stall force. PACS numbers: Valid PACS appear here Active behavior of skeletal muscles is associated with time scales of about 30 ms [1]. At shorter times (∼1 ms) muscles exhibit a nontrivial passive response: if a tetanized muscle is suddenly extended, it comes loose, and if it is shortened, it tightens up with apparently no involvement of Adenosine Triphosphate (ATP) [2]. This unusual mechanical behavior, associated with the unfold- ing of the attached myosin cross-linkers (cross-bridges), qualifies muscles as metamaterials [3]. As we argue be- low, an important factor in this behavior is the domi- nance of parallel connections with multiple shared links entailing cooperative effects; see Fig. 1. Similar mean- field coupling can be found in many hierarchical biolog- ical systems [4]; in particular, it plays a crucial role in cell adhesion, where individual binding elements interact through a common elastic background [5]. Interaction-induced synchronization during muscle contractions reveals itself through macroscopic fluctua- tions and spatial inhomogeneities [6]. In ratchet-based and chemomechanical models such collective behavior is usually attributed to breaking of the detailed balance [7], with long range interactions entering the problem implic- itly as a force dependence of the chemical rates [8]. How- ever, the cooperative behavior of myosin cross-bridges can be detected during short time force recovery [9], and therefore the origin of synchronization should be within reach of models disregarding disequilibrium and activ- ity. In this Letter, we show that already equilibrium response of "muscle material" is associated with highly synchronized behavior at the microscale which explains its unusual passive response. In particular, we show that an order-disorder phase transition is displayed by the celebrated Huxley-Simmons (HS) model [2] if, instead of physiological isometric load- ing conditions (length clamp) also known as a hard de- vice, one considers isotonic (load clamp) loading condi- tions or a soft device. While a considerable difference in behavior of a muscle loaded in these two different ways can be deduced from experimental data [10], the origin of the disparity has been so far unexplained. We argue that behind it is a nonequivalence of equilibrium ensem- bles ubiquitous in systems with long range interactions [11]. Most remarkably, we find that a careful parameter fit- ting places the actual skeletal muscle almost exactly into a ferromagnetic Curie point. This agrees with the ob- servation [2] that the effective stiffness of skeletal mus- cles associated with fast force recovery is close to zero in the state of isometric contractions and strongly suggests that muscles are finely tuned to perform near marginal stability. Other experimentally observed manifestations of criticality include kinetic slowing down and large scale macroscopic fluctuations near the stall force conditions [9]. We demonstrate the robustness of our predictions by comparing the HS model, which we interpret as a hard spin description, with a regularized (RHS) model where filament elasticity is taken into consideration and con- ventional spins are replaced by elastic snap springs. FIG. 1. Schematic structure of the three layers of organization inside a sarcomere: (a) global architecture with domineering parallel links; (b) structure of an elementary contractile unit shown in more detail in Fig. 2; (c) individual attached cross- bridge represented by a bistable element in series with a shear spring. The HS model. -- We consider a prototypical model of a half-sarcomere with N attached cross-bridges arranged in parallel [2]; at time scales of fast force recovery N can be considered constant [2, 10]. Each cross-bridge is modeled as a series connection of a bistable spin unit and a linear (shear) spring; see Fig. 2. We use dimensionless variables with the power-stroke size a as a unit of displacement and κa2 as the unit of energy, where κ is the stiffness of the series spring. Then the spin variable takes values xi = 0 (pre-power-stroke state) and xi = −1 (post-power-stroke 3 1 0 2 n u J 7 2 ] h p - o i b . s c i s y h p [ 2 v 4 9 0 5 . 5 0 3 1 : v i X r a (a)(b)(c) 2 where the "regular solution" term is responsible for co- operative (ferromagnetic) behavior. In Fig. 3, we show the position of the minima of g(p) when t is chosen to ensure that in the paramagnetic phase p(t, β) = 1/2. In the disordered (high temperature) state all cross-bridges are in random conformations, while in the ordered (low temperature) state the system exhibits coherent fluctuations between post-power-stroke and pre- power-stroke configurations. These fluctuations describe temporal microstructures responsible for the plateau in the force-elongation relation z = t − p, where p is a solu- tion of t − p + 1/2 − v0 + (1/β) ln[p/(1 − p)] = 0. FIG. 3. Bifurcation diagram for the HS model placed in a soft device showing synchronized states (A and C) and disordered state (B). Solid lines shows minima of the Gibbs free energy at t = 1. The critical point is located at β = 4. The equilibrium Gibbs energy is concave because g(cid:48)(cid:48)(t) = −1 − βN(cid:104)(p − p)2(cid:105) ≤ 0, so in the soft device the stiffness is always positive. Since in the hard de- vice the stiffness is sign indefinite, the two ensembles are not equivalent. This is expected for systems with strong long range interactions that are inherently nonadditive [11, 12]. Negative stiffness in the hard device HS model has been known for a long time [2, 8, 13, 14]; however, it was not previously associated with the particular internal architecture of muscle material. As we have already mentioned, the original HS fit of experimental data [2] places the system exactly into the critical state (Curie point). In this state the cor- relation length diverges and fluctuations become macro- scopic, which is consistent with observations at stall force conditions [6, 15]. This suggests that skeletal muscles, as many other biological systems, may be tuned to criti- cality. The proximity to the critical point would then be the result of either evolutionary or functional self- organization. The marginal stability of the critical state allows the system to amplify interactions, ensure strong feedback, and achieve considerable robustness in front of random perturbations. In particular, it is a way to quickly switch back and forth between highly efficient synchronized stroke and stiff behavior in the desynchro- nized state. FIG. 2. Elementary contractile unit. (a) Energy of the bistable (power-stroke) element: HS model (thin line) and RHS model (thick line); (b) N cross-bridges in hard device. In the HS model, κ0, κ1, κf = ∞ and y = z. state), and the total energy per particle in the hard device v0 is the energetic bias of the pre-power-stroke state. In this formulation, the HS model describes the simplest paramagnetic spin system. is v(x, z) = (1/N )(cid:80)[(1 + xi)v0 + (1/2) (z − xi)2], where cle f (z, β) = −[1/(N β)] ln(cid:82) exp [−βN v(x, z)] dx where β = κa2/(kbθ). At fixed p = −(1/N )(cid:80) xi, represent- At finite temperature θ, the equilibrium behavior of this system is characterized by the free energy per parti- ing the fraction of cross-bridges in the post-power-stroke state, the macroscopic ( N → ∞ ) free energy takes the form f = p (z + 1)2 + (1 − p) z2 + v0 + 1 β S(p) (cid:21) (cid:20) 1 2 (cid:21) (cid:20) 1 2 (cid:26) (cid:20) (cid:21) where S(p) = p log(p) + (1 − p) log(1 − p). The function f (p) is always convex with a minimum at p(z, β) = 1/2− (1/2) tanh[(β/2)(z − v0 + 1/2)]. The equilibrium tension per cross-bridge is then t = (z + p), which is exactly the formula found by Huxley and Simmons. The equilibrium free energy is (cid:20) (cid:18) z2 (cid:19)(cid:21)(cid:27) 1 β β 2 − 2 + v0 ln exp −β + exp (z + 1)2 f = − and the susceptibility (stiffness) is t(cid:48)(z) = f(cid:48)(cid:48)(z). The function f (z) is convex for β < 4 and is nonconvex for β > 4 exhibiting a range with negative equilibrium stiff- ness (metamaterial behavior). If we take the values from [2], a = 8 nm and κa2/2 = 2kbθ, we obtain β = 4, which corresponds to zero stiffness at the state of isometric con- tractions z0 = v0 − 1/2. In the soft device setting, not studied by Hux- ley and Simmons, the energy becomes w(x, z, t) = v(x, z) − tz, where t is the applied force per parti- cle. Now the variable z plays the role of an inter- nal parameter whose adiabatic elimination produces a Curie-Weiss mean-field potential depending on ((cid:80) xi)2. −[1/(N β)] ln(cid:82) exp [−βN w(x, z, t)] dxdz. At fixed p we The equilibrium Gibbs free energy is now g(t, β) = obtain in the thermodynamic limit g = − 1 2 t2 + pt + (1 − p)v0 + 1 2 p(1 − p) + 1 β S(p), x0uHS−a(a)κ1κ0uRHSκ1,κ0κNκfyz(b)0246800.20.40.60.81ACBβp0101Bpg0101ABCpgA:B:C: 3 where λ1 = κ1/κ, λ0 = κ0/κ. Second, we introduce a mixed device (mimicking myofilament elasticity [7, 16]) by adding to our parallel bundle a series spring. The resulting energy per cross-bridge in a hard device is N(cid:88) i=1 (cid:2)uRHS(xi) + (y − xi)2(cid:3) + 1 2 λf 2 (z − y)2, v(x,y;z) = 1 N FIG. 4. Recovery rates in hard and soft devices. Symbols: Postprocessing of experimental data; see [10]. Open symbols, hard device; filled symbols, soft device. Dashed lines: HS model in hard (h) and soft (s) devices; parameters are taken from [2]. Solid lines: RHS model in hard (h) and soft (s) devices obtained from stochastic simulations; parameters have been fit to experimental data: λ1 = 0.41, λ0 = 1.21, λf = 0.72, l = −0.08, N = 112, β = 52 (κ = 2 pN/nm, a = 10 nm, θ = 277.13 K), z0 = 4.2 nm/hs. The ensembles nonequivalence in the HS model has also a kinetic signature. Experiments on quick recovery reveal that muscle fibers react to load steps much slower than to length steps [2, 10]. This agrees with our model, where coherent response (in isotonic conditions) is always slower than disordered response (in isometric conditions). Indeed, by using Kramers approximation Huxley and Simmons obtained in a hard device the kinetic equation (cid:104)p(cid:105) = −k−(cid:104)p(cid:105) + k+(1 − (cid:104)p(cid:105)), where (cid:104)p(cid:105) is the average over ensemble. The constants k+, k− satisfy the detailed balance k+/k− = exp [−β(z − v0 + 1/2)], and the recov- ery rate is 1/τ = k−{1 + exp[−β(z − v0 + 1/2)]}. In a soft device we may use the same model with z = t − (cid:104)p(cid:105), which accounts for force-dependent chemistry and intro- duces nonlinear feedback. The characteristic rate around a given state (cid:104)p(cid:105) is then 1/τ = k−{1 + [1 − β(1 − (cid:104)p(cid:105))] exp[−β(t − (cid:104)p(cid:105) − v0 + 1/2)]}. When (cid:104)p(cid:105) is small, t − (cid:104)p(cid:105) > z, and the relaxation in a soft device is slower than in a hard device. In Fig. 4, we show the rates ob- tained from the HS model; in the case of a soft device, the nonlinear kinetic equation was solved numerically for the duration 10 ms. We see that in a soft device the rates are indeed slower than in a hard device, however, the experimental measurements are not matched quanti- tatively. The RHS model. -- To test the robustness of the HS mechanism of synchronization and to achieve quantita- tive agreement with kinetic data, we now consider a nat- ural regularization of the HS model. First, following [14] we replace hard spins by soft spins described by a piece- wise quadratic double well potential -- see Fig. 2(a) -- (cid:40) 1 uRHS(x) = 2 λ0(x)2 + v0 2 λ1(x + 1)2 1 if x > l, if x ≤ l, where y is a new internal variable and λf = κf /(N κ); see Fig. 2(b). It is clear that our lump description of filament elasticity misrepresents short range interactions [17]; however, this should not affect our main results [18]. To study the soft device case we must again consider the energy w(x, y, z, t) = v(x, y, z) − tz, where t is the applied force per cross-bridge. A transition from hard to soft ensemble is made by taking the limit λf → 0, z → ∞ with λf z → t. At finite λf the RHS model can be viewed as a version of the mean-field ϕ4 model studied in [4, 11]. The HS model is a limiting case of the RHS model with λ1,0 → ∞ and λf → ∞. The first of these lim- its allows one to replace continuous dynamics by jumps and use the language of chemical kinetics; however, it also erases information about the barriers; see [14]. The second limit eliminates the Curie-Weiss (mean-field) in- teraction among individual cross-bridges at fixed z, and that is why the synchronized behavior was overlooked in [2]. FIG. 5. Phase diagram for the RHS model in a hard device with z selected to ensure that (cid:104)p(cid:105) = 1/2 at each point (β, λf ). In the shaded region, the function f (p) is nonconvex which leads to coherent fluctuations. Outside this region, fluctu- ations are not synchronized. The cross indicates an almost critical configuration with realistic parameters (used in Fig. 4). Equilibrium behavior in the RHS model can be again described analytically, because it is just a redressed HS model. In the limit λf → ∞ the function f (p) is convex as in HS model, while at finite values of λf it is now non- convex. This shows that in the RHS model the account of filament elasticity brings about phase transition (and bistability) also in the hard device. The bistable nature of the macroscopic free energy in both soft and hard devices implies that the system can −15−10−5001234HS(s)HS(h)RHS(h)RHS(s)δz=z−z0(nm/hs)rate(1/ms)Hard[2]Soft[10]×01002000123βλf0101time(s)p01pf0101time(s)p01pf01pf be in two coherent states, and therefore within a large set of half-sarcomeres one should expect observable spatial inhomogeneities. This prediction is in agreement with ubiquitous "off-center" displacements of M lines recorded during isometric contractions [6]. The phase diagram showing the role of filament elastic- ity in hard device is shown in Fig. 5. The dependence of the critical temperature on λf suggests that actomyosin systems can control the degree of cooperativity by tun- ing the internal stiffness; likewise, variable stiffness of the loading device may be used in experiments to either ac- tivate or deactivate the collective behavior. Notice that the realistic choice of parameters again selects a near crit- ical state; the exact criticality is compromised since the symmetry between the pre- and post-power-stroke states is now broken (as λ1 (cid:54)= λ0) and the phase transition be- comes weakly first order; see Fig.6. A behavior similar to our synchronization has been also observed in the models of passive adhesive clusters, where the elastic feedback appears as strain- or force-dependent chemistry [5]. Given that the two systems exhibit almost identical cooperative behavior, we expect criticality to be also a factor in the operation of focal adhesions. FIG. 6. Bifurcation diagram for nonsymmetric RHS model with realistic parameters: (a) hard device with z = 0.37; (b) soft device with t = 0.21; this loading secures that (cid:104)p(cid:105) = 1/2 for β = 52. Parameters are as in Fig. 4. Inset (a) corresponds to β = 52, inset (b) to β = 25. The two ensembles, soft and hard, remain inequiv- alent in the RHS model. Thus, in the soft device the equilibrium Gibbs free energy g is concave since g(cid:48)(cid:48)(t) = −1/λf − βN(cid:104)(y − (cid:104)y(cid:105))2(cid:105) ≤ 0 which means that the stiffness is always positive. Instead in the hard de- vice f(cid:48)(cid:48)(z) = λf [1 − βN(cid:104)(y − (cid:104)y(cid:105))2(cid:105)], and the stiffness can be both positive and negative. While negative stiff- ness should be a characteristic feature of realistic half- sarcomeres (see Fig. 7), it has not been observed in ex- periments on whole myofibrils. The reason may be that in myofibrils a single half-sarcomere is never loaded in a hard device. The effective dimensionless temperature may also be higher because of the quenched disorder, and the stiffness may be smaller due to nonlinear elasticity. One can also expect the unstable half-sarcomeres to be stabilized actively through processes involving ATP hy- drolysis [19]. 4 To study kinetics in the RHS model we perform di- rect numerical simulations by using a Langevin ther- mostat. We assume that the macroscopic variables y and z are fast and are always mechanically equilibrated which is not an essential assumption. The response of the remaining variables xi is governed by the system dxi = b(xi)dt +(cid:112)2β−1dBi, where the drift is b(x, z) = −u(cid:48)RHS(xi)+(1+λf )−1(λf z +N−1(cid:80) xi)−xi in a hard de- vice and b(x, t) = −u(cid:48)RHS(xi) + t + N−1(cid:80) xi − xi in a soft device. In both cases the diffusion term dBi represents a standard Wiener process. FIG. 7. (a) States attainable during quick recovery: solid line, T2; dotted line, L2; dashed line, Lqs 2 corresponds to quasis- tationary states; dash-dotted line, T1 and L1. Symbols show experimental points for hard (open) and soft (filled) devices; see [10]. In a hard device, the equilibrium T2 curve coincides with the results of stochastic simulations. In a soft device, the equilibrium L2 curve differs from the simulation results at 10 ms (Lqs 2 curve). Averaged trajectories after abrupt loading at 1 ms: (b) in a hard device and (c) in a soft device. Curve (1): δz = −1 nm/hs; curve (2): δz = −5 nm/hs; curve (3): t/t0 = 0.9; curve (4): t/t0 = 0.5 with t0 = 0.21. Other parameters are as in Fig.4. In Fig. 7, we show the results of stochastic simulations imitating quick recovery experiments [2]. The system, initially in thermal equilibrium at fixed z0 (or t0), was perturbed by applying fast (∼ 100 µs) length (or load) steps with various amplitudes. In a soft device the sys- tem was not able to reach equilibrium within the exper- imental time scale. Instead, it remained trapped in a quasistationary (glassy) state because of the high energy barrier associated with collective power stroke. Such ki- netic trapping which fits the pattern of two-stage dynam- ics exhibited by systems with strong long range interac- tions [11, 20] may explain the failure to reach equilib- rium in experiments reported in [9]. In the hard device case, the cooperation among the cross-bridges is weaker and kinetics is much faster, allowing the system to reach equilibrium at the experimental time scale. A quantita- tive comparison of the rates obtained in our simulations with experimental values (see Fig. 4) shows that the RHS model reproduces the kinetic data in both hard and soft ensembles rather well. In conclusion, we mention that the prototypical na- 02040608010000.20.40.60.81βp020406080100β01pf01pg(a)(b)−10−500.511.5T1,L1T2=f0(z)L2=−g0(t)Lqs2δz(nm/hs)t/t0Hard[2]Soft[10](a)0.51(1)(2)t/t0(b)024−10−50(3)(4)time(ms)δz(c) ture of our model implies that passive collective behavior should be a property common to general cross-linked ac- tomyosin networks. We have shown that the degree of cooperativity in such networks can be strongly affected by elastic stiffness of the filaments. This suggests that a generic system of this type can be tuned to criticality by an actively generated prestress [21]. We thank D. Chapelle, J.-F. Joanny, K.Kruse, and V. Lombardi for insightful comments. M. Caruel thanks Monge PhD fellowship from Ecole Polytechnique for fi- nancial support. ∗ [email protected] [1] B. Alberts, A. Johnson, J. Lewis, M. Raff, K. Roberts, and P. Walter, Molecular Biology of the Cell, 5th ed. (Garland Science, New York, 2007); J. Howard, Mechan- ics of Motor Proteins and the Cytoskeleton, edited by Sinauer (Sinauer, Sunderland, MA, 2001). [2] A. F. Huxley and R. M. Simmons, Nature 233, 533 (1971); L. E. Ford, A. F. Huxley, and R. M. Simmons, J. Physiol. 269, 441 (1977); V. Lombardi, G. Piazzesi, and M. Linari, Nature 355, 638 (1992); G. Piazzesi, F. Francini, M. Linari, and V. Lombardi, J. Physiol. 445, 659 (1992); M. Linari, G. Piazzesi, and V. Lom- bardi, Biophys. J. 96, 583 (2009). [3] Z. G. Nicolaou and A. E. Motter, Nature Mater. 11, 608 (2012). [4] K. Kometani and H. Shimizu, J. Stat. Phys. 13, 473 (1975); R. C. Desai and R. Zwanzig, ibid. 19, 1 (1978); T. Dauxois, S. Lepri, and S. Ruffo, Commun. Nonlinear Sci. 8, 375 (2003). [5] T. Erdmann and U. S. Schwarz, Eur. Phys. J. E 22, 123 (2007). [6] P. Y. Pla¸cais, M. Balland, T. Gu´erin, J. F. Joanny, and P. Martin, Phys. Rev. Lett. 103, 158102 (2009); I. Pavlov, R. Novinger, and D. E. Rassier, J. Biomech. 42, 2808 (2009); Y. Shimamoto, M. Suzuki, S. V. Mikhailenko, K. Yasuda, and S. Ishiwata, Proc. Natl. Acad. Sci. U.S.A. 106, 11954 (2009); E. L. F. Holzbaur and Y. E. Goldman, Curr. Opin. Cell Biol. 22, 4 (2010); S. Ishiwata, Y. Shimamoto, and N. Fukuda, Prog. Bio- phys. Molec. Biol. 105, 187 (2011). [7] F. Julicher and J. Prost, Phys. Rev. Lett. 75, 2618 (1995); F. Julicher, A. Ajdari, and J. Prost, Rev. Mod. Phys. 69, 1269 (1997); T. Guerin, J. Prost, P. Martin, 5 and J.-F. Joanny, Curr. Opin. Cell Biol. 22, 14 (2010); T. Guerin, J. Prost, and J. F. Joanny, Eur. Phys. J. E 34, 60 (2011). [8] T. Duke, Proc. Natl. Acad. Sci. U.S.A. 96, 2770 (1999); Philos. T. Roy. Soc. B 355, 529 (2000); D. Smith, M. Geeves, J. Sleep, and S. Mijailovich, Ann. Biomed. Eng. 36, 1624 (2008); T. Erdmann and U. Schwarz, Phys. Rev. Lett. 108, 188101 (2012). [9] C. F. Armstrong, A. F. Huxley, and F. J. Julian, J. Phys- iol. 186, 26P (1966); K. A. P. Edman, 404, 301 (1988); H. Granzier, A. Mattiazzi, and G. Pollack, Am. J. Phys- iol. 259, C266 (1990); K. A. P. Edman and N. Curtin, J. Physiol. - London 534, 553 (2001). [10] G. Piazzesi, L. Lucii, and V. Lombardi, J. Physiol. 545, 145 (2002); M. Reconditi, M. Linari, L. Lucii, A. Stew- art, Y. Sun, P. Boesecke, T. Narayanan, R. Fischetti, T. Irving, G. Piazzesi, M. Irving, and V. Lombardi, Na- ture 428, 578 (2004); G. Piazzesi et.al., Cell 131, 784 (2007). [11] A. Campa, T. Dauxois, and S. Ruffo, Physics Reports 480, 57 (2009). [12] A. Campa, S. Ruffo, and H. Touchette, Physica A 385, 233 (2007); R. S. Ellis, J. Stat. Phys. 101, 999 (2000). [13] T. L. Hill, Prog. Biophys. Molec. Biol. 28, 267 (1974); 29, 105 (1976). [14] L. Marcucci and L. Truskinovsky, Phys. Rev. E 81, 051915 (2010); Eur. Phys. J. E 32, 411 (2010). [15] I. Telley, J. Denoth, E. Stussi, G. Pfitzer, and R. Stehle, Biophys. J. 90, 514 (2006); I. Pavlov, R. Novinger, and D. E. Rassier, Am J Physiol-Cell Ph 297, C1211 (2009); T. Serizawa, T. Terui, T. Kagemoto, A. Mizuno, T. Shi- mozawa, F. Kobirumaki, S. Ishiwata, S. Kurihara, and N. Fukuda, 301, C1116 (2011). [16] H. Huxley, A. Stewart, H. Sosa, and T. Irving, Bio- phys. J. 67, 2411 (1994); K. Wakabayashi, Y. Sugimoto, H. Tanake, Y. Ueno, Y. Takezawa, and Y. Amemiya, 67, 2422 (1994). [17] L. E. Ford, A. F. Huxley, and R. M. Simmons, J. Phys- iol. 311, 219 (1981); S. Mijailovich, J. Fredberg, and J. Butler, Biophys. J. 71, 1475 (1996); P.-G. de Gennes, C.R. Acad. Sci. IV-Phys. 2, 1505 (2001). [18] J. Nagle, Phys. Rev. A 2, 2124 (1970); M. Kardar, Phys. Rev. B 28, 244 (1983). [19] A. Vilfan and E. Frey, J. Phys.: Condens. Matter 17, S3901 (2005). [20] F. Bouchet, S. Gupta, and D. Mukamel, Physica A 389, 4389 (2010); C. Nardini, S. Gupta, S. Ruffo, T. Dauxois, and F. Bouchet, J. Stat. Mech. 2012, L01002 (2012). [21] M. Sheinman, C. P. Broedersz, and F. C. MacKintosh, Phys. Rev. Lett. 109, 238101 (2012).
1705.03033
1
1705
2017-05-08T18:22:12
Towards a Field Theoretical Stochastic Model for Description of Tumour Growth
[ "physics.bio-ph", "q-bio.TO" ]
We develop a field theory-inspired stochastic model for description of tumour growth based on an analogy with an SI epidemic model, where the susceptible individuals (S) would represent the healthy cells and the infected ones (I), the cancer cells. From this model, we obtain a curve describing the tumour volume as a function of time, which can be compared to available experimental data.
physics.bio-ph
physics
Towards a Field Theoretical Stochastic Model for Description of Tumour Growth Leonardo Mondaini∗ Department of Physics, Federal University of the State of Rio de Janeiro - UNIRIO, Rio de Janeiro, Brazil Abstract We develop a field theory-inspired stochastic model for description of tumour growth based on an analogy with an SI epidemic model, where the susceptible individuals (S) would represent the healthy cells and the infected ones (I), the cancer cells. From this model, we obtain a curve de- scribing the tumour volume as a function of time, which can be compared to available experimental data. Keywords: Second Quantization Approach, Stochastic Epidemic Models, Master Equation, Tumour Growth, Liver Cancer 7 1 0 2 y a M 8 ] h p - o i b . s c i s y h p [ 1 v 3 3 0 3 0 . 5 0 7 1 : v i X r a ∗Electronic address: [email protected] 1 I. INTRODUCTION In a recent letter [1], we have shown how the standard field theoretical language based on creation and annihilation operators (building blocks of the second quantization method [2]) may be used for a straightforward derivation of closed master equations [3] describing the population dynamics of multivariate stochastic epidemic models. This was mainly motivated by the observation that, as remarked in [4], for the kinds of model studied in population biology and epidemiology, a field theoretical description is notationally neater and more manageable than standard methods, in often replacing sets of equations with single equa- tions with the same content. Indeed, a single hamiltonian function sums up the dynamics compactly, even when births and deaths allow the population size to change, and may be easily written down from a verbal description of the transitions presented in these models. In the present work we employ the very same methodology established in [1] to develop a field-theory inspired stochastic model for description of tumour growth based on an analogy with an SI epidemic model [5], where the susceptible individuals (S) would represent the healthy cells and the infected ones (I), the cancer cells. From this model, we were able to obtain a curve describing the time evolution of the tumour volume, which is then compared to available experimental data for liver cancer. Our main motivation comes from the pioneering ideas about cancer as a phase transition presented in [6], more specifically, from the following observations: (1) progression of cancer must also involve a population-level shift due to a competition between two co-existing phenotypes: normal and cancerous, and (2) given enough time and resources, cancer cells will usually outcompete healthy cells in the organ or tissue where they coexist, in the competiton for space and resources. The rest of this work is organized as follows. In Section 2, we introduce the basic aspects of our model which allow us to obtain differential equations describing the time evolution of the mean number of individuals in the interacting populations we are dealing with, i.e., normal (healthy) and cancer cells. The analytical solutions for these differential equations, besides the specific curve describing the time evolution of tumour volume and its comparison to experimental data are presented in Section 3. Finally, in Section 4 we present our concluding remarks. 2 II. BUILDING THE MODEL We will start by considering interacting populations, whose total sizes are allowed to change, composed of two types of individuals: the normal (healthy) cells and the tumour (cancer) cells. Let us introduce N (t) and C(t) as random variables which represent, respec- tively, the number of normal and cancer cells at a given time instant t. We will then consider a bivariate process {(N (t); C(t))}∞ t=0 with a joint probability func- tion given by (1) Our aim is to compute time-dependent expectation values of the observables N (t) and p(n,m)(t) = Prob{N (t) = n; C(t) = m}. C(t), which may be defined in terms of the configuration probability according to (cid:104)N (t)(cid:105) = (cid:104)C(t)(cid:105) = n p(n,m)(t); m p(n,m)(t). Let us represent the probabilistic state of the system by the vector with the normalization condition(cid:80) n,m p(n,m)(t) = 1. µ, ν(cid:105) = p(n,m)(t)n, m(cid:105) , As an example, the vector 1 4 (1, 1(cid:105) + 2, 1(cid:105) + 1, 2(cid:105) + 2, 2(cid:105)) represents the probability dis- tribution where there are 1 or 2 healthy/cancer cells present, each one with probability 1/4, i.e. p(1,1) = p(2,1) = p(1,2) = p(2,2) = 1 4. Since the configurations are given entirely in terms of occupation numbers (n, m), which calls for a representation in terms of second-quantized bosonic operators [7], we will introduce creation and annihilation operators for the normal cells, respectively, h† and h, and for the cancer cells, namely, c† and c. These operators must satisfy the following commutation relations (cid:2)h, h†(cid:3) =(cid:2)c, c†(cid:3) = 1; [h, c] =(cid:2)h, c†(cid:3) =(cid:2)c, h†(cid:3) =(cid:2)h†, c†(cid:3) = 0. (cid:88) (cid:88) n,m n,m (cid:88) n,m 3 (2) (3) (4) TABLE I: Transition rates presented in our model and corresponding terms in the hamiltonian H. Description Transition N bh−→ N + N birth of normal cell (rate bh) C bc−→ C + C birth of cancer cell (rate bc) N dh−→ ∅ C dc−→ ∅ N λ−→ C death of normal cell (rate dh) Contribution to H bh(h†h − h†h†h) = bh(n − h†n) bc(c†c − c†c†c) = bc(m − c†m) dh(h†h − h) = dh(n − h) dc(c†c − c) = dc(m − c) death of cancer cell (rate dc) change normal → cancer (rate λ) λ(h†h − c†h) = λ(n − c†h) As usual in the second quantization framework, we say that h† and c† "create" , respectively, normal and cancer cells when applied over the reference (vacuum) state 0, 0(cid:105). This allows us to build our space from basis vectors of the form n, m(cid:105) =(cid:0)h†(cid:1)n(cid:0)c†(cid:1)m 0, 0(cid:105). This vacuum state has the following properties: h0, 0(cid:105) = c0, 0(cid:105) = 0 (from which "anni- hilation" operators) and (cid:104)0, 00, 0(cid:105) = 1 (inner product). Following the above definitions, we also have h† n, m(cid:105) = n + 1, m(cid:105) ; h n, m(cid:105) = nn − 1, m(cid:105) ; c† n, m(cid:105) = n, m + 1(cid:105) ; c n, m(cid:105) = mn, m − 1(cid:105) . (5) At this point it is worth to note that h†h n, m(cid:105) = nn, m(cid:105) and c†c n, m(cid:105) = mn, m(cid:105). Thus, the operators n = h†h and m = c†c just count the number of cells in a definite state. This is the main reason why they are usually called number operators. The vector state of our system may be then rewritten in terms of creation and annihilation operators as (cid:88) p(n,m)(t)(cid:0)h†(cid:1)n(cid:0)c†(cid:1)m 0, 0(cid:105) . µ, ν(cid:105) = (6) The time evolution of our system will then be generated by a linear operator H (called hamiltonian) which may be constructed directly from the transition rates present in our n,m model according to Table I (cf. [4], Table 1). Note that, upon summing up the terms presented in Table I, we may write our hamiltonian as H = + (bh + dh + λ) n + (bc + dc) m −(cid:0)bhh†n + bcc†m + dhh + dcc + λc†h(cid:1) . (7) The notational advantage of this field theoretical description is made clear at this point if we observe that, from the above definitions, the equation which represents the flux of 4 probability between states at rates defined by our model (the so-called master equation or forward Kolmogorov differential equation [5]) takes the very compact form of an imaginary- time Schrodinger-type linear equation, namely µ, ν(cid:105) = −H µ, ν(cid:105) . d dt (8) We can get, after some algebra, a more common representation for the master equation by substituting the expressions for the hamiltonian (7) and the vector state (3) into Equation (8) d dt p(n,m) = −[(bh + dh + λ)n + (bc + dc)m] p(n,m) +bh(n − 1) p(n−1,m) + bc(m − 1) p(n,m−1) +dh(n + 1) p(n+1,m) + dc(m + 1) p(n,m+1) (9) In order to compute the time-dependent expectation values of the observables N (t) and C(t) through the above master equation, we will follow the well-established methodology presented in [5] and introduce the following moment generating function (mgf ) +λ(n + 1) p(n+1,m−1). p(n,m)enφ+mθ. (10) Note that from the above equation we have After multiplying Equation (9) by exp(nφ + mθ) and summing on (n, m), we are led, (cid:88) n,m M (φ, θ; t) =(cid:10)eφN (t)eθC(t)(cid:11) = (cid:21) (cid:20)∂M (cid:20)∂M (cid:21) =(cid:10)Ck(t)(cid:11) ; (cid:88) (cid:88) (cid:21) φ,θ=0 φ,θ=0 ∂φ ∂θ n,m n,m = = n p(n,m) = (cid:104)N (t)(cid:105) ; m p(n,m) = (cid:104)C(t)(cid:105) (cid:20) ∂kM ∂φk (cid:21) φ,θ=0 =(cid:10)N k(t)(cid:11) . (cid:88) (cid:0)e−φ − 1(cid:1) + λ(cid:0)e−φ+θ − 1(cid:1)(cid:3) ∂M (cid:0)eφ − 1(cid:1) + dh (cid:0)eθ − 1(cid:1) + dc (cid:0)e−θ − 1(cid:1)(cid:3) ∂M +(cid:2)bc ∂M ∂t d p(n,m) enφ+mθ n,m=0 ∂φ dt = . ∂θ 5 and, in general (cid:20)∂kM ∂θk φ,θ=0 after some algebra, to = +(cid:2)bh (11) (12) (13) Finally, by differentiating the above equation with respect to φ and evaluating the result at φ = θ = 0 we get the following differential equation for (cid:104)N (t)(cid:105) = d dt (cid:104)N (t)(cid:105) = (bh − dh − λ)(cid:104)N (t)(cid:105) . (14) (cid:20) ∂2M (cid:21) ∂t ∂φ φ,θ=0 (cid:20) ∂2M (cid:21) ∂t ∂θ φ,θ=0 On the other hand, if we differentiate Equation (13) with respect to θ and evaluate at φ = θ = 0 we get the following differential equation for (cid:104)C(t)(cid:105) = d dt (cid:104)C(t)(cid:105) = λ(cid:104)N (t)(cid:105) + (bc − dc)(cid:104)C(t)(cid:105) . (15) III. COMPARISON TO LIVER CANCER DATA. FITTING THE PARAMETERS IN THE ANALYTICAL SOLUTION FOR (cid:104)C(t)(cid:105) By defining (cid:104)N (0)(cid:105) ≡ N0, (cid:104)C(0)(cid:105) ≡ C0, βh ≡ bh − dh and βc ≡ bc − dc we obtain the following analytical solutions for Equations (14) and (15) (cid:104)N (t)(cid:105) = N0 e(βh−λ)t (16) and (cid:2)e(βh−λ)t − eβct(cid:3) + C0eβct. (cid:104)C(t)(cid:105) = λN0 (17) Finally, if we consider that the number of normal cells is approximately constant (βh ≈ 0) and that the volume of a cancer cell (v) is approximately the same of a normal cell, we may βh − λ − βc write the following expression for the time evolution of the tumour volume (cid:2)e−λt − eβct(cid:3) + (Vc)0eβct, λ(Vh)0 −λ − βc Vc(t) = (18) where Vc(t) ≡ v (cid:104)C(t)(cid:105), (Vh)0 ≡ vN0 is the initial volume of the normal tissue, and (Vc)0 ≡ vC0 is the initial tumour volume. The above expression is compared to experimental data for liver cancer (average tumour volume) in Figure 1. The corresponding data have been obtained through the analysis of computed tomography (CT) scans for a set of 34 patients available in [8]. IV. CONCLUDING REMARKS We would like to finish this work by making a few comments about Equation (18). As far as we know, this is the first time that an expression describing the time evolution of the 6 FIG. 1: The solid line represents Equation (18) for an initial total volume of the patient liver (VT )0 = (Vh)0 + (Vc)0 = 2153.00 cm3; λ = −7.07 × 10−3 months−1; βc = 5.67 × 10−1 months−1 and (Vc)0 = 29.87 cm3. This curve was obtained by fitting the experimental data available in [8] (square points in the figure), using the software OriginPro 8 [9]. volume of tumours [11] is obtained from basic assumptions about cancer as a phase transition and, most certainly, this is the very first model for description of tumour growth built by using standard field theoretical language commonly found in models describing fundamental interactions of elementary particles. In a future work we are going to present a qualitative analysis of the behaviour of solutions for the system of first order linear differential equations composed by (14) and (15) in the phase plane, which we believe will shed more light on how our model is indeed connected to the idea of cancer as a phase transition. Last but not least, other possible extensions for the present model should consider the inclusion of other kinds of dependence in the birth/death rates of cells, as temperature and/or concentration gradients of toxic carcinogens, for example. Acknowledgments The author would like to thank Prof. Jack A. Tuszynski for the valuable discussions and the very stimulating scientific environment shared at University of Alberta's Li Ka Shing 7 Applied Virology Institute, where the main results of this work were obtained. [1] Mondaini, L. (2015) Second Quantization Approach to Stochastic Epidemic Models. Biomed. Sci. Today, 2, e8. http://www.mdtcanada.ca/bmst/v2/e8.pdf [2] Greiner, W. and Reinhardt, J. (1996) Field Quantization. Springer-Verlag, Berlin. [3] Gardiner, C. W. (2004) Handbook of Stochastic Methods for Physics, Chemistry and the Natural Sciences, 3rd Ed. Springer, Heidelberg. [4] Dodd, P. J. and Ferguson, N. M. (2009) A Many-Body Field Theory Ap- proach to Stochastic Models in Population Biology. PLoS ONE, 4, e6855. http://dx.doi.org/10.1371/journal.pone.0006855 [5] Allen, L. J. S. (2011) An Introduction to Stochastic Processes with Applications to Biology. Chapman and Hall/CRC, Boca Raton. [6] Davies, P. C. W., Demetrius, L. and Tuszynski, J. A. (2011) Cancer as a dynamical phase transition. Theor. Biol. Med. Model., 8, 30. http://www.tbiomed.com/content/8/1/30 [7] Cardy, J. L. and Tauber, U. C. (1998) Field Theory of Branching and Annihilating Random Walks. J. Stat. Phys., 90, 1. http://dx.doi.org/10.1023/A:1023233431588 [8] Lieffers, J. R. et al. (2009) A viscerally driven cachexia syndrome in patients with advanced colorectal cancer: contributions of organ and tumor mass to whole-body energy demands. Am. J. Clin. Nutr., 89, 1173. http://dx.doi.org/10.3945/ajcn.2008.27273 [9] OriginLab Corp. (2007) OriginPro 8. OriginLab Corp., Northampton. [10] Benzekry, S. et al. (2014) Classical Mathematical Models for Description and Prediction of Experimental Tumor Growth. PLoS Comput. Biol., 10, e1003800. http://dx.doi.org/10.1371/journal.pcbi.1003800 [11] A comprehensive review of the most common mathematical models for description of tumour growth may be found in [10]. 8
1808.03746
1
1808
2018-08-11T03:24:23
A New Boron-10 Delivery Agent for Boron Neutron Capture Therapy: Fluorescent Boron-10 Embedded Nanodiamonds
[ "physics.bio-ph", "physics.med-ph" ]
Boron neutron capture therapy is a powerful anti-cancer treatment, the success of which depends heavily on the boron delivery agent. Enabling the real-time tracing of delivery agents as they move through the body is crucial to the further development of boron neutron therapy. In this study, we fabricate highly bio-compatible boron-10 embedded nanodiamonds using physical ion implantation in conjunction with a two-step annealing process. The red fluorescence of the nanodiamonds allows their use in fluorescence microscopy and in vivo imaging systems, thereby making it possible to conduct tracking in real time. The proposed fluorescent boron-10 embedded nanodiamonds, combining optical visibility and boron-10 transport capability, are a promising boron delivery agent suitable for a wide range of biomedical applications.
physics.bio-ph
physics
A New Boron-10 Delivery Agent for Boron Neutron Capture Therapy: Fluorescent Boron-10 Embedded Nanodiamonds Bo-Rong Lin1, Srinivasu Kunuku2, Chien-Hsu Chen2, Tzung-Yuang Chen3, Tung-Yuan Hsiao2, Yu-Jen Chang4, Li-Chuan Liao4, Huan Niu2* and Chien-Ping Lee1 1 Department of Electronics Engineering and Institute of Electronics, National Chiao Tung University, Hsinchu, Taiwan 2Accelerator Laboratory, Nuclear Science and Technology Development Center, National Tsing Hua University, Hsinchu, Taiwan 3Health Physics Division, Nuclear Science and Technology Development Center, National Tsing Hua University, Hsinchu, Taiwan 4Bioresource Collection and Research Center, Food Industry Research and Development Institute, Hsinchu, Taiwan Electronic mail: [email protected] Boron neutron capture therapy is a powerful anti-cancer treatment, the success of which depends heavily on the boron delivery agent. Enabling the real-time tracing of delivery agents as they move through the body is crucial to the further development of boron neutron therapy. In this study, we fabricate highly bio-compatible boron-10 embedded nanodiamonds using physical ion implantation in conjunction with a two-step annealing process. The red fluorescence of the nanodiamonds allows their use in fluorescence microscopy and in vivo imaging systems, thereby making it possible to conduct tracking in real time. The proposed fluorescent boron-10 embedded nanodiamonds, combining optical visibility and boron-10 transport capability, are a promising boron delivery agent suitable for a wide range of biomedical applications. 1 Radiation therapy1), chemotherapy2), and surgery3) are the three most common approaches to cancer treatment. Radiation therapy has undergone considerable advances in recent decades, particularly in the fields of photon therapy4) and particle therapy, involving protons5), neutrons6), and heavy particles7). Boron neutron capture therapy (BNCT) involves the use of heavy particles as an alternative to conventional radiation therapy for the treatment of tenacious cancers, such as those of the head and neck cancer8-10) and glioblastoma multiforme11, 12) (GBM). BNCT is based on the nuclear capture and fission reactions that occur when a non- radioactive boron-10 absorbs a neutron13, 14). This form of nuclear fission produces high- energy alpha particles and high-energy lithium-7, both of which induce ionization in the immediate vicinity of the reaction. The range of ionization is 4 -- 9 µm, which is less than the diameter of a tumor cells. After the boron-10 ions are internalized by tumor cells, high-energy particles kill the tumor cells by damaging their DNA or organelles14-16). Many factors can affect the outcomes of BNCT treatment, particularly the delivery agents used for the transport of boron-10 to tumor tissue. Boron delivery agents must be low in toxicity17). They must permit the retention of boron-10 in the cancer tissue, while promoting rapid clearance from the body17). Finally, they must be traceable throughout the body18). Overall, an ideal delivery agent would have the ability to target cancer cells with large numbers of boron-10 atoms, while remaining traceable and nontoxic to the human body. High efficiency boron delivery agent is still under developing. One of the options is using nanoparticles to achieve those points. Nanoparticles have been used with considerable success in the diagnosis and treatment of cancer19). They are often combined with conventional cancer treatment drugs to enable the direct targeting of tumor tissue, thereby reducing the effects of toxicity on the immune system and normal tissue20). Nanodiamonds (NDs) provide excellent bio-compatibility21-24) and their surface is easily modified to facilitate the attachment of conventional drugs that kill tumor cells 2 via chemical processes25). Most previous studies in this field have focused exclusively on the surface of NDs (i.e., disregarding the interior of the NDs) for drug delivery or other features. Recently, we developed a novel technique that uses ion implantation to enable the addition of boron-10 to NDs for use as a boron delivery agent. Atoms implanted within NDs are stable and provide good bio-compatibility26). This technique allows us to add a new dimension for BNCT applications. In this work, we fabricate boron-10 embedded nanodiamonds using physical ion implantation in conjunction with a two-step annealing process. During the fabrication process, high-energy boron ions penetrate into the diamond lattice to create vacancies. Annealing is then used to form nitrogen-vacancy (NV) centers within the NDs (first step) and to remove graphite from the surface (second step). It is generally not possible to observe NDs after they enter cells or animal body; however, the proposed boron-10 embedded NDs with red fluorescence can be observed using fluorescence microscopy or an in vivo imaging system (IVIS). With optical visibility, boron-10 embedded NDs become an observable boron delivery agent. One has the chance to observe them by red fluorescence with low background signals in this wavelength region27). Boron-10 embedded NDs overcome the limitation that the existing boron delivery agents are not easy to be observed inside the cells or animal body. Furthermore, once combining with targeting ability by NDs surface modification28), boron- 10 embedded NDs can become akin to a kind of Trojan Horse. The Trojan people (cells) will pull the horses into their city (the interior of cells). In other words, boron-10 embedded NDs will more accumulate in tumor and create higher therapeutic efficiency. It will solve the long existing problem that current boron delivery agents have weak targeting performance in the last century. To summarize, fluorescent boron-10 embedded NDs provide a novel boron delivery agent to extend the practical use of NDs and open up new avenues of research for the further development of BNCT. 3 The NDs powder with average 100 nm in diameter from Microdiamant Co. was dissolved in DI water and then the solution was coated on silicon wafer. Boron-10 ions were implanted into NDs with energy of 30 keV, dose of 3x1015 ions/cm2 and energy 60 keV, dose of 3x1015 ions/cm2. After implantation, the wafer was cut to many pieces to perform different processes. The annealing processes were performed at 800°C for 2 hours under vacuum and 450°C for 3 hours under air using Nabertherm annealing furnace. X-ray photoelectron spectroscopy(XPS) spectra were obtained using PHI Quantera II system. Microscope fluorescence images were obtained by Nikon Ti-U inverted research fluorescence microscopy. The excitation wavelength range is 560(cid:3399)12.5nm and the detected emission wavelength range is 692(cid:3399)40nm. In vivo imaging system IVIS Lumina II was used to capture IVIS fluorescence images. The excitation wavelength is 535 nm and the detected emission wavelength is 680 nm. Following boron-10 ion implantation, our first objective was to confirm that the implanted NDs contained boron ions. Fig. 1 presents the B 1s XPS spectra of as-implanted NDs and NDs. The characteristic peak of boron29) near 190 eV can be clearly observed in Fig. 1 (a). The measured spectra revealed that the as-implanted NDs contained boron ions and we called them B-NDs. For practical bio-applications, it is crucial that the nanoparticles used for tracing provide fluorescence of a sufficient intensity to allow observation using a microscope or other device. Many nitrogen atoms are included in the ND structure during fabrication. Boron-10 ion implantation produces many vacancies in the diamond lattice. In this study, we used high- temperature vacuum annealing to promote the formation of NV centers to facilitate red fluorescence. Initial annealing under vacuum at 800°C was followed by air annealing at 450°C for the removal of graphite layers, which might otherwise block the fluorescence. Two-step annealing methods such as this have previously been used to repair the lattice of NDs and remove surface graphite30). We used XPS spectra to evaluate the effects of the two-step annealing process on B-NDs. All obtained XPS spectra were fitted with possible carbon chemical states31, 32). Figs. 2 (a) 4 and (b) respectively present the C 1s spectra of the original (non-implanted) NDs and as- implanted B-NDs. The difference of calculated sp3/sp2 ratio between the original NDs and as- implanted B-NDs is insignificant. It implies the damage caused by boron-10 ion implantation. Fig. 2 (c) presents the C 1s spectra of B-NDs annealed under vacuum at 800°C. The calculated sp3/sp2 ratio of the B-NDs increased from 2.74 before annealing to approximately 7 after annealing. This is a clear demonstration of how high-temperature vacuum annealing repairs the lattice structure of as-implanted B-NDs. Fig. 3 (d) presents the C 1s spectrum of B-NDs following two-step annealing. The carbide signal likely originated at the contact surface between the NDs and the silicon substrate. The calculated sp3/sp2 ratio of the B-NDs increases from 7 prior to second-stage annealing to 11.6 after this step. This clearly demonstrates the efficacy of air annealing at 450°C for the removal of graphite from the surface of the B-NDs. To summarize, first-stage annealing formed NV centers, and second-stage annealing removed surface graphite, thereby allowing the tracking of red fluorescence. In this study, we employed a microscope commonly used to observe fluorescence in biology labs. Fig. 3 presents fluorescence images of NDs and B-NDs obtained using the microscope under various conditions. Only the B-NDs that underwent the proposed two-step annealing process (Fig. 3 (d)) are clearly observable. Thus, it would be reasonable to expect that even after the B-NDs are internalized by cells, they could be observed in a similar manner. We also used non-invasive IVIS system to verify our findings. The IVIS system uses the full wavelength spectrum from high-energy blue to low-energy infrared to study disease progression, cell trafficking, and gene expression patterns via fluorescence or luminescence33). The detection of B-NDs using the IVIS system is a clear indication of the efficacy of these B-NDs in practical bio-applications. The emission wavelength of NV centers of maximum intensity is approximately 680 nm34). We used a halogen lamp light source with a wavelength of 535 nm to excite the samples and receive fluorescence emission signals with a wavelength of 680 nm. Fig. 4 presents IVIS fluorescence images of the original NDs and B-NDs under various 5 conditions. The grey image is taken by visible light camera and then the fluorescent signal is overlaid on this grey image. Only the colored area presented fluorescence emissions at 680 nm. Again, the only visible fluorescence was from B-NDs after two-step annealing. Testing was also conducted on NDs that underwent the proposed two-step annealing process without boron- 10 implantation. Those NDs remained invisible under the instruments used in this study. Visibility under a fluorescence microscope and the IVIS system provides compelling evidence of the practical value of the proposed B-NDs. In conclusion, this paper reports the fabrication of fluorescent boron-10 embedded NDs using boron-10 ion implantation followed by a two-step annealing process. The B-NDs annealed at 800°C for 2 hours under vacuum and 450°C for 3 hours under air provided red fluorescence of sufficient intensity to enable detection using a commercial fluorescence microscope as well as the IVIS system. The proposed B-NDs, combining optical visibility and boron-10 transport capability, are a promising boron delivery agent suitable for a wide range of biomedical applications. Acknowledgement The authors acknowledge Team Union Ltd. for their financial supporting. The authors thank Dr. Causon K.C. Jen at Axcelis Technologies., Inc. for his valuable inputs. 6 S. Sotirios: Medical Physics. 37 [3](2010)1374. I. Tannock: Cancer Treat Rep. 62 [8](1978)1117. [in eng]. R.A. Patchell, P.A. Tibbs, J.W. Walsh, R.J. Dempsey, Y. Maruyama, R.J. Kryscio, W.R. Markesbery, J.S. Macdonald and B. Young: New England Journal of Medicine. 322 [8](1990)494. S.S. Korreman: Phys Med Biol. 57 [23](2012)R161. R. Mohan and D. Grosshans: Adv Drug Deliv Rev. 109 (2017)26. G.E. Laramore, J.M. Krall, F.J. Thomas, K.J. Russell, M.H. Maor, F.R. Hendrickson, K.L. Martz and T.W. Griffin: Am.J.Clin.Oncol. 16 [2](1993)164. T. Ohno: EPMA J. 4 [1](2013)9. I. Kato, K. Ono, Y. Sakurai, M. Ohmae, A. Maruhashi, Y. Imahori, M. Kirihata, M. Nakazawa and Y. Yura: Applied Radiation and Isotopes. 61 [5](2004)1069. L.W. Wang, S.J. Wang, P.Y. Chu, C.Y. Ho, S.H. Jiang, Y.W.H. Liu, Y.H. Liu, H.M. Liu, J.J. Peir, F.I. Chou, S.H. Yen, Y.L. Lee, C.W. Chang, C.S. Liu, Y.W. Chen and K. Ono: Applied Radiation and Isotopes. 69 [12](2011)1803. R.F. Barth, M. H Vicente, O.K. Harling, W.S. Kiger, K.J. Riley, P.J. Binns, F.M. Wagner, M. Suzuki, T. Aihara, I. Kato and S. Kawabata: Radiation Oncology. 7 [1](2012)146. Reference 1) 2) 3) 4) 5) 6) 7) 8) 9) 10) 20) 21) 11) M. Chadha, J. Capala, J.A. Coderre, E.H. Elowitz, J.-i. Iwai, D.D. Joel, H.B. Liu, L. 12) Wielopolski and A.D. Chanana: International Journal of Radiation Oncology*Biology*Physics. 40 [4](1998)829. R. Henriksson, J. Capala, A. Michanek, S.-Å. Lindahl, L.G. Salford, L. Franzén, E. Blomquist, J.-E. Westlin and A.T. Bergenheim: Radiotherapy and Oncology. 88 [2](2008)183. R.F. Barth, A.H. Soloway and R.G. Fairchild: Cancer Research. 50 [4](1990)1061. S.J. Walker: Radiography. 4 [3](1998)211. 13) H.J. Taylor and M. Goldhaber: Nature. 135 (1935)341. 14) 15) 16) Y. Yura and Y. Fujita: Oral Science International. 10 [1](2013)9. 17) P.J. Kueffer, C.A. Maitz, A.A. Khan, S.A. Schuster, N.I. Shlyakhtina, S.S. Jalisatgi, J.D. Brockman, D.W. Nigg and M.F. Hawthorne: Proc Natl Acad Sci U S A. 110 [16](2013)6512. 18) M.F. Hawthorne and M.W. Lee: Journal of Neuro-Oncology. 62 [1](2003)33. 19) S.D. Steichen, M. Caldorera-Moore and N.A. Peppas: Eur J Pharm Sci. 48 [3](2013)416. S. Tran, P.J. DeGiovanni, B. Piel and P. Rai: Clin Transl Med. 6 [1](2017)44. R.J. Narayan, W. Wei, C. Jin, M. Andara, A. Agarwal, R.A. Gerhardt, C.-C. Shih, C.- 7 M. Shih, S.-J. Lin, Y.-Y. Su, R. Ramamurti and R.N. Singh: Diamond and Related Materials. 15 [11-12](2006)1935. 22) Y. Yuan, Y. Chen, J.-H. Liu, H. Wang and Y. Liu: Diamond and Related Materials. 18 [1](2009)95. 23) Y. Zhu, J. Li, W. Li, Y. Zhang, X. Yang, N. Chen, Y. Sun, Y. Zhao, C. Fan and Q. 24) Huang: Theranostics. 2 [3](2012)302. L. Moore, J. Yang, T.T. Lan, E. Osawa, D.K. Lee, W.D. Johnson, J. Xi, E.K. Chow and D. Ho: ACS Nano. 10 [8](2016)7385. 25) X. Wang, X.C. Low, W. Hou, L.N. Abdullah, T.B. Toh, M. Mohd Abdul Rashid, D. 26) Ho and E.K.-H. Chow: ACS Nano. 8 [12](2014)12151. B.R. Lin, C.H. Chen, S. Kunuku, T.Y. Chen, T.Y. Hsiao, H. Niu and C.P. Lee: Sci Rep. 8 [1](2018)7058. 27) A.M. Smith, M.C. Mancini and S. Nie: Nat Nanotechnol. 4 [11](2009)710. 28) B.-M. Chang, H.-H. Lin, L.-J. Su, W.-D. Lin, R.-J. Lin, Y.-K. Tzeng, R.T. Lee, Y.C. Lee, A.L. Yu and H.-C. Chang: Advanced Functional Materials. 23 [46](2013)5737. E.A. Il'inchik, V.V. Volkov and L.N. Mazalov: Journal of Structural Chemistry. 46 [3](2005)523. S. Osswald, G. Yushin, V. Mochalin, S.O. Kucheyev and Y. Gogotsi: Journal of the American Chemical Society. 128 [35](2006)11635. P. Mérel, M. Tabbal, M. Chaker, S. Moisa and J. Margot: Applied Surface Science. 136 [1](1998)105. F. Le Normand, J. Hommet, T. Szörényi, C. Fuchs and E. Fogarassy: Physical Review B. 64 [23](2001)235416. 29) 30) 31) 32) 33) N. Andreu, A. Zelmer and S. Wiles: FEMS Microbiology Reviews. 35 [2](2011)360. 34) A.M. Zaitsev: Optical properties of diamond: a data handbook (Springer Science & Business Media, 2013). 8 Figure caption Fig. 1. B 1s XPS spectra of (a) as-implanted NDs. (b) original NDs. Fig. 2. C 1s XPS spectra of (a) Original NDs. (b) as-implanted B-NDs. (c) B-NDs annealed at 800°C under vacuum. (d) B-NDs annealed at 800°C under vacuum and 450°C in air. Fig. 3. Microscope fluorescence images of NDs and B-NDs under different condition. (a) Original NDs. (b) As implanted B-NDs (c) B-NDs annealed at 800°C under vacuum. (d) B- NDs annealed at 800°C under vacuum and 450°C in air. Scale bar: 10 μm. Fig. 4. IVIS fluorescence images of NDs and B-NDs under different condition. (a) Original NDs. (b) As implanted B-NDs (c) B-NDs annealed at 800°C under vacuum. (d) B-NDs annealed at 800°C under vacuum and 450°C in air. The grey image is taken by visible light camera and then the fluorescent signal is overlaid on this grey image. Only the colored area presented fluorescence emissions at 680 nm. 9 Fig. 1. 10 Fig. 2. 11 Fig. 3. 12 Fig. 4. 13
1803.06751
1
1803
2018-03-18T21:45:12
Plasmonic Toroidal Metamolecules Assembled by DNA Origami
[ "physics.bio-ph", "cond-mat.soft" ]
We demonstrate hierarchical assembly of plasmonic toroidal metamolecules, which exhibit tailored optical activity in the visible spectral range. Each metamolecule consists of four identical origami-templated helical building blocks. Such toroidal metamolecules show stronger chiroptical response than monomers and dimers of the helical building blocks. Enantiomers of the plasmonic structures yield opposite circular dichroism spectra. The experimental results agree well with the theoretical simulations. We also demonstrate that given the circular symmetry of the structures, distinct chiroptical response along their axial orientation can be uncovered via simple spin-coating of the metamolecules on substrates. Our work provides a new strategy to create plasmonic chiral platforms with sophisticated nanoscale architectures for potential applications such as chiral sensing using chemically-based assembly systems.
physics.bio-ph
physics
Maximilian J. Urban,1,† Palash K. Dutta,1,‡ Pengfei Wang,1,‡ Xiaoyang Duan,† Xibo Shen,† Baoquan Ding,*, Yonggang Ke,*,‡ Na Liu,*,†,§ † Max Planck Institute for Intelligent Systems, Heisenbergstrasse 3, D-70569 Stuttgart, Germany ‡ Wallace H. Coulter Department of Biomedical Engineering, Georgia Institute of Technology and Emory University, Atlan- ta, Georgia 30322, United States CAS Key Laboratory of Nanosystems and Hierarchical Fabrication, CAS Center for Excellence in Nanoscience, National Center for Nanoscience and Technology, Beijing 100190, China § Kirchhoff Institute for Physics, University of Heidelberg, Im Neuenheimer Feld 227, D-69120, Heidelberg, Germany Supporting Information Placeholder ABSTRACT: We demonstrate hierarchical assembly of plas- monic toroidal metamolecules, which exhibit tailored optical activity in the visible spectral range. Each metamolecule consists of four identical origami-templated helical building blocks. Such toroidal metamolecules show stronger chiroptical response than monomers and dimers of the helical building blocks. Enantiomers of the plasmonic structures yield opposite circular dichroism spectra. The experimental results agree well with the theoretical simulations. We also demonstrate that given the circular sym- metry of the structures, distinct chiroptical response along their axial orientation can be uncovered via simple spin-coating of the metamolecules on substrates. Our work provides a new strategy to create plasmonic chiral platforms with sophisticated nanoscale architectures for potential applications such as chiral sensing using chemically-based assembly systems. Realization of three-dimensional (3D) plasmonic nanoarchitec- tures exemplifies one of the key challenges in nanotechnology.1 In general, there are two basic types of nanofabrication strategies: top-down and bottom-up. Top-down approaches such as electron- beam lithography and focused-ion beam etching offer custom patterns with relatively low throughput and resolution.2 The fabri- cated 3D plasmonic nanostructures are generally fashioned in layers and spaced with dielectrics. This posts many challenges for downstream applications. In contrast, bottom-up self-assembly approaches,3 which often involve molecular recognition charac- ters, offer large-scale fabrication, biochemical compatibility, and many other benefits. Among a variety of materials for self- assembly, DNA represents one of the most attractive building blocks largely due to its unprecedented programmability. DNA has been used to construct increasingly complex structures4 and to pattern nanoparticles.5 In particular, the DNA origami4 technique allows for creation of fully addressable DNA templates with custom-designed shapes and sizes. DNA origami enables rational 3D organization of metal nanoparticles with nanometer precision, allowing for engineering complex plasmonic architectures with tailored optical response. Recently, plasmonic chiral nanomaterials have gained increas- ingly attention due to its potential applications in nanophotonics, ultrasensitive sensing, and etc.6 Natural chiral molecules such as proteins and DNA only exhibit weak optical activity in the UV range. Plasmonic assemblies composed of metal nanoparticles that are arranged in chiral configurations offer an elegant way to achieving strong optical activity in the visible spectral range.6 For example, in the seminal work of Liedl et al, plasmonic helices were realized by assembling gold nanoparticles (AuNPs) on ori- gami bundles in a staircase fashion7 Nevertheless, previous work has mostly focused on studies of fundamental plasmonic chiral building blocks (e.g. helices7 and tetramers8), and averaged chi- roptical response over all structural orientations, because the assemblies were randomly oriented in aqueous solutions. In this communication, we demonstrate the first experimental realization of plasmonic toroidal metamolecules through hierar- chical assembly of helical building blocks templated by DNA origami. Each structure consists of four plasmonic helical mono- mers arranged in a toroidal geometry.9 Systematic investigations Figure 1. (a) A curved DNA origami monomer that comprises 24 DNA helices in a bundle is formed by a scaffold strand and numerous staple strands. (b) Four origami monomers form an origami ring through head- head and tail-tail connectors. (c) Schematic of the left-handed (LH) and right-handed (RH) 3D plasmonic toroidal metamolecules. Each meta- molecule has a toroidal diameter of 120 nm and consists of 24 AuNPs (13 nm). Figure 2. (a) Left: Agarose gel electrophoresis images of the origami monomers, dimers, and rings. Right: Agarose gel electrophoresis images of the plas- monic LH helical monomers, dimers, and toroidal metamolecules. (b) TEM image of the origami rings (scale bar 200 nm). (c) TEM image of the LH plas- monic helical monomers (scale bar 200 nm). (d) TEM image of the LH plasmonic helical dimers (scale bar 200 nm). TEM images of the LH (e) and RH (f) plasmonic toroidal metamolecules (scale bar 400 nm). The frame size of each inset image is 200 nm. on different experimental conditions are carried out to achieve high-quality plasmonic structures for tailored optical functionali- ties. The assembled metamolecules with designated handedness exhibit pronounced optical activity in the visible spectral range. The experimental optical spectra agree well with the theoretical predictions. We also demonstrate that given the unique circular symmetry, distinct chiroptical response along the axial orientation of the toroidal metamolecules can be revealed by simple spin- coating of the structures on substrates. Such orientation self- alignment neither requires surface functionalization nor introduc- es birefringence effects, which often arise from symmetry break- ing. This is of great importance for achieving reliable chiroptical results on the single structure level using dark-field spectroscopy as well as very practical for development of polarization conver- sion devices using chemically-based assembly systems. In addi- tion, the axial chiroptical response can be enhanced through silver (Ag) coating of the AuNP-based structures. Elemental mapping with energy-filtered transmission electron microscopy (TEM) clearly demonstrates the formation of Au core/Ag shell nanoparti- cles in a toroidal geometry. The plasmonic toroidal metamolecules are assembled utilizing DNA origami technique, as illustrated in Fig. 1. A long single- stranded DNA scaffold (M13 bacteriophage DNA) is folded by ~180 short staple strands through hybridization to form an origa- mi monomer (Figs. 1a, S1, and S2; also see supporting infor- mation for more design details). Each monomer contains 24 curved DNA helices bundled in a honeycomb lattice.10 Four mon- omers are connected together using 12 head-head and 12 tail-tail connector-strands to form a complete origami ring of ~120 nm in terms of inner-diameter (Figs. 1b and S3). Half-ring structures, i.e., dimers, can be achieved by linking two monomers using either the head-head or tail-tail connectors. To assemble AuNPs on origami, six binding sites arranged either in left-handed (LH) or right-handed (RH) fashion are extended from each monomer. Each binding site contains three capture strands of the same se- quence. Overall, 24 binding sites are distributed evenly along one complete origami ring. To ensure that each AuNP binds only to one site, two sets of capture strands with different sequences are used for the binding sites on the monomers in an alternating fash- ion. AuNPs (13 nm in diameter) functionalized with complemen- tary DNA strands are assembled onto the origami ring to form a LH or RH plasmonic toroidal metamolecule (Fig. 1c). Previously, Shih et al. have demonstrated a gear-like origami structure of ~100 nm in diameter using four 18-helix homo- monomers but the yield was lower than 10%.11 To ensure a high yield of our origami rings, which is critical to create high-quality plasmonic structures for strong chiroptical response, careful opti- mization of the ring design and assembly protocol have been carried out. First, we have tested two versions (V1 and V2) of monomer designs with different degrees of curvatures (Figs. S1- S7). Mean angles of the V1 and V2 monomers are 99±9° and 79±12°, respectively (Fig. S7). The V1 monomers show higher yield for forming origami rings, while the V2 monomers give rise to mainly larger concatemers. Hence, subsequent optimization has been carried out using the V1 origami. Second, the influence from the sticky-end strength of the con- nectors on the hierarchical assembly12 of the origami rings is investigated. Each connector comprises two segments. One seg- ment is anchored on the edge of one monomer and the other seg- ment that contains 12 sticky-ends (either 1-base or 3-base long) is anchored on the edge of another monomer. Systematic studies using gel electrophoresis and TEM clearly show that connectors with 3-base long sticky-ends give rise to higher yield than those with 1-base long sticky-ends (Fig. S8a). Stronger sticky-ends with higher binding forces may minimize non-specific binding and thus facilitate the formation of the origami rings. Figure 3. (a) Stepwise formation of the plasmonic toroidal metamolecules. (b) Schematic for the CD measurement of the plasmonic nanostructures that are dispersed in water with random orientations. Measured (c) and simulated (d) CD spectra of the plasmonic helical monomers, dimers, and toroidal meta- molecules. (e) Charge distribution (Ez) of the plasmonic structure at resonance. (f) CD components of the LH plasmonic toroidal metamolecules along different incident directions of the circularly polarized light. The averaged CD response is presented by the black curve. A comparison between the rings formed either directly or in a stepwise manner from purified monomers and the rings formed from purified dimers shows no substantial difference in yield (Fig. S8b). At last, the effects of the dimer concentration, assembly time, and connector-strand concentration on the assembly yield of the origami rings are examined (Figs. S9 and S10). Experimental results show that lower dimer concentrations (0.4 and 0.2 nM) lead to significantly improved yield (~49% and ~67%, respective- ly) of the origami rings, when compared to 32% from a higher dimer concentration of 1 nM. The incorrect assemblies are mostly aggregated spiral-like structures as observed from gel electropho- resis and TEM images (Fig. S9). Lower assembly concentrations are associated with slower assembly rates and lower possibilities of aggregation, which could both contribute to higher yield of the origami rings. In another assay experiment on various assembly times, many incomplete rings are observed, when the incubation time is less than 6 hours (Fig. S10a). Structure analysis using gel electrophoresis shows that a minimal 50-fold excess of the con- nectors is required to achieve high yield of the origami rings (Fig. S10b). Utilizing the optimized design and assembly protocol, DNA-origami monomers, dimers, and rings are successfully assembled with high yield, as confirmed by gel electrophoresis (Fig. 2a, left panel) and TEM imaging (Fig. 2b). The inner- diameter of the rings is measured to be 120.0±3.9 nm based TEM images (Fig. S11). Subsequently, plasmonic helical monomers, dimers, as well as toroidal metamolecules in LH and RH geome- tries are assembled via attaching AuNPs onto the origami precur- sors. AuNPs functionalized with single-stranded DNA are mixed with the corresponding purified origami at a 10-fold excess of particles per binding site. Both gel electrophoresis (Fig. 2a, right panel) and TEM images (Figs. 2c-2f) confirm the successful assembly of the plasmonic nanostructures (Fig. S12). The plas- monic toroidal structures were purified by gel electrophoresis prior to TEM imaging (Fig 2e, f). Particle numbers fewer than 24 are observed from the individual plasmonic structures. This is likely due to steric hindrance between the AuNPs as well as due to the inner-capture strands that are less accessible than the outer ones (Fig. S13). When interacting with light, plasmons excited in the AuNPs can be collectively coupled.6 The helical arrangement of the AuNPs imprints a "corkscrew" character to the collective plas- mons, introducing a twist of specific handedness into the propaga- tion direction of light, which leads to different absorption in re- sponse to LH and RH circularly polarized light, i.e., circular dichroism (CD). Figure 4. (a) Schematic for the axial CD component measurement of the plasmonic toroidal metamolecules. Circularly polarized light is perpendic- ularly incident on the structures. (b) Left: SEM image of the toroidal metamolecules after Ag enhancement. Right: Energy-filtered TEM images of the Ag-enhanced metamolecules. Au and Ag are represented by the blue and yellow colors, respectively. Frame size: 205 nm (c) Measured and simulated CD spectra of the LH and RH samples after one (top), three (middle), and ten (bottom) spin-coating cycles of the respective structures on the glass substrates. The structure density after ten spin coating cycles increased to approximately ~0.15 ring/m2, which is also utilized in the simulations. In order to characterize the chiroptical response of the plas- monic helical monomers, dimers, and toroidal metamolecules, CD measurements have been carried out using a Jasco-1500 CD spectrometer. To keep the AuNP concentration in all measure- ments consistent, the CD spectra are measured from the same solution by first addition of the head-head connectors into mono- mers to form dimers and subsequent addition of the tail-tail con- nectors to form toroidal metamolecules (Figs. 3a and 3b). The LH monomers, dimers, and toroidal metamolecules exhibit character- istic peak-to-dip line shapes with increasing amplitudes, while their RH counterparts display mirrored spectra correspondingly. The plasmonic toroidal metamolecules show the most pronounced CD spectra, when compared to the constituent plasmonic helical monomers and dimers (Figs. 3c and S14). The experimental re- sults agree well with the simulated results (Fig. 3d). The charge distribution at resonance is presented in Fig. 3e, in which the dipole coupling nature between the neighboring AuNPs can be clearly identified. In the aqueous solution, the plasmonic meta- molecules are randomly oriented and illuminated by light of fixed incident direction (Fig. 3b). In simulations, this is equivalent to averaging over the CD components along all possible incident directions. Fig. 3f presents the simulated CD components along different directions, as well as the averaged CD response of the LH plasmonic toroidal metamolecules. Next, the CD response along the axial orientation of the toroidal metamolecules is explored by spin-coating the respective LH and RH structures on glass substrates after silver enhancement. The incident direction of light is perpendicular to the substrate and therefore the axial CD component of the structures becomes accessible (Fig. 4a). Silver enhancement is utilized in order to achieve enhanced CD signals using a relatively small amount of the structures (Fig. S15). The deposition of Ag shells on the AuNPs in the assemblies is verified by scanning electron micros- copy (SEM) imaging and elemental-mapping imaging using energy-filtered TEM (Fig. 4b). A 1.5 L of the LH or RH sample is first spin-coated on a glass substrate at 4000 rpm, which yields a very weak CD response (Fig. 4c, top panel), owing to a low structure density of < 0.1 rings/m2 (Fig. S16). Through repeti- tive spin-coatings, the structure density is eventually increased to ~0.15 ring/m2 (bottom panel). As a result, the CD response becomes increasingly stronger. The LH and RH samples exhibit bisignate spectral profiles, which are nearly mirrored (Fig. 4c). The experimental spectra show broader spectral profiles when compared to the simulated spectra (Fig. 4c) because of inhomoge- neous broadening mainly resulting from the Ag shell thickness variations over different AuNPs. In contrast, a uniform Ag shell (5 nm) over different AuNPs is utilized in the simulations. The spin coating and drying process may also cause distortion to the toroidal geometry. The elimination of birefringence effects is confirmed by measuring all the samples from both the front and back incident directions (Fig. S17). Simulations on toroidal dimer structures and fused toroidal structures can be found in Figs. S18 and S19, respectively. In conclusion, we have demonstrated DNA origami-templated plasmonic toroidal metamolecules, which yield tailored optical activity in the visible spectral range. These plasmonic toroidal metamolecules show stronger chiroptical response than the con- stituent plasmonic chiral building blocks such as helical mono- mers and dimers. These toroidal metamolecules can be dispersed on solid substrates for generating chiroptical response along their axial orientation. This might offer a new pathway to create plas- monic platforms with tunable CD for enantiomer sensing. The CD intensity dependence on the structure density could also be useful for development of polarization conversion devices, which allow for specific polarizations of light with controlled intensities. Detailed experimental procedures, origami design, DNA sequenc- es, gel and TEM images, and structure density analysis after spincoating. This material is available free of charge via the Inter- net at http://pubs.acs.org. *[email protected];*[email protected]; *[email protected] 1: M.J.U.; P.K.D. and P.W. contributed equally to this work. The authors declare no competing financial interests. We thank K. Hahn for assistance with elemental mapping and HR-TEM microscopy. TEM data was collected at the Stuttgart Center for Electron Microscopy (StEM). We thank Thomas Weiss and Anna Asenjo-Garcia for initial calculations. This research was supported by the Sofja Kovalevskaja grant from the Alexander von Humboldt-Foundation, the Marie Curie CIG grant, and the European Research Council (ERC Dynamic Nano) grant. We also thank the support from the Wallace H. Coulter Department of Biomedical Engineering Faculty Startup Grant and a Winship Cancer Institute Billi and Bernie Marcus Research Award. (1) Liu, Y.; Zhang, X. Chem. Soc. Rev. 2011, 40, 2494-2507. (2) Soukoulis, C. M.; Wegener, M. Nature Photon. 2011, 5, 523-530. (3) Jones, M. R.; Osberg, K. D.; Macfarlane, R. J.; Langille, M. R.; Mirkin, C. A. Chem. Rev. 2011, 111, 3736-3827. (4) Rothemund, P. W. Nature 2006, 440, 297-302. (5) (a) Pinheiro, A. V.; Han, D.; Shih, W. M.; Yan, H. Nat. Nanotechnol. 2011, 6, 763-772; (b) Torring, T.; Voigt, N. V.; Nangreave, J.; Yan, H.; Gothelf, K. V. Chem. Soc. Rev. 2011, 40, 5636-5646; (c) Li, Y.; Liu, Z.; Yu, G.; Jiang, W.; Mao, C. J. Am. Chem. Soc. 2015, 137, 4320-4323; (d) Tian, Y.; Wang, T.; Liu, W.; Xin, H. L.; Li, H.; Ke, Y.; Shih, W. M.; Gang, O.; Nat. Nanotechnol. 2015, 10, 637-644; (e) Lan, X.; Lu, X.; Shen, C.; Ke, Y.; Ni, W.; Wang, Q.; J. Am. Chem. Soc. 2015, 137, 457-462. (f) Sharma, J.; Chhabra, R.; Liu, Y.; Ke, Y.; Yan, H. Angew. Chem. Int. Ed. 2006, 45, 730-735. (6) (a) Fan, Z.; Govorov, A. O. Nano Lett. 2010, 10, 2580-2587. (b) Zhou, C.; Duan, X.; Liu, N. Nat. Commun. 2015, 6, 8102; (c) Urban, M. J.; Zhou, C.; Duan, X.; Liu, N. Nano Lett. 2015, 15, 8392-8396. (7) Kuzyk, A.; Schreiber, R.; Fan, Z.; Pardatscher, G.; Roller, E. M.; Hogele, A.; Simmel, F. C.; Govorov, A. O.; Liedl, T. Nature 2012, 483, 311-314. (8) (a) Shen, X.; Asenjo-Garcia, A.; Liu, Q.; Jiang, Q.; Garcia de Abajo, F. J.; Liu, N.; Ding, B. Nano Lett. 2013, 13, 2128-2133; (b) Mastroianni, A. J.; Claridge, S. A.; Alivisatos, A. P. J. Am. Chem. Soc. 2009, 131, 8455-8459. (9) Kaelberer, T.; Fedotov, V. A.; Papasimakis, N.; Tsai, D. P.; Zheludev, N. I. Science 2010, 330, 1510-1512. (10) Douglas, S. M.; Marblestone, A. H.; Teerapittayanon, S.; Vazquez, A.; Church, G. M.; Shih, W. M. Nucleic Acids Res. 2009, 37, 5001-5006. (11) Dietz, H.; Douglas, S. M.; Shih, W. M. Science 2009, 325, 725- 730. (12) (a) Liu, W.; Zhong, H.; Wang, R.; Seeman, N. C. Angew.Chem. Int. Ed. 2011, 50, 264-267; (b) Li, Z.; Liu, M.; Wang, L.; Nangreave, J.; Yan, H.; Liu, Y. J. Am. Chem. Soc. 2010, 132, 13545-13552; (c) Rajendran, A.; Endo, M.; Katsuda, Y.; Hidaka, K.; Sugiyama, H. ACS Nano 2011, 5, 665-671; (d) Iinuma, R.; Ke, Y.; Jungmann, R.; Schlichthaerle, T.; Woehrstein, J. B.; Yin, P. Science 2014, 344, 65-69; (e) Jungmann, R.; Scheible, M.; Kuzyk, A.; Pardatscher, G.; Castro, C. E.; Simmel, F. C. Nanotechnology 2011, 22, 275301; (f) He, Y.; Ye, T.; Su, M.; Zhang, C.; Ribbe, A. E.; Jiang, W.; Mao, C. Nature 2008, 452, 198- 201.
1807.11765
1
1807
2018-07-31T11:26:03
Phototactic Robot Tunable by Sensorial Delays
[ "physics.bio-ph", "cond-mat.soft" ]
The presence of a delay between sensing and reacting to a signal can determine the long-term behavior of autonomous agents whose motion is intrinsically noisy. In a previous work [M. Mijalkov, A. McDaniel, J. Wehr, and G. Volpe, Phys. Rev. X 6, 011008 (2016)], we have shown that sensorial delay can alter the drift and the position probability distribution of an autonomous agent whose speed depends on the illumination intensity it measures. Here, using theory, simulations, and experiments with a phototactic robot, we generalize this effect to an agent for which both speed and rotational diffusion depend on the illumination intensity and are subject to two independent sensorial delays. We show that both the drift and the probability distribution are influenced by the presence of these sensorial delays. In particular, the radial drift may have positive as well as negative sign, and the position probability distribution peaks in different regions depending on the delay. Furthermore, the presence of multiple sensorial delays permits us to explore the role of the interaction between them.
physics.bio-ph
physics
Phototactic Robot Tunable by Sensorial Delays Maximilian Leyman,1, 2, ∗ Freddie Ogemark,1, 2, ∗ Jan Wehr,3 and Giovanni Volpe1 1Department of Physics, University of Gothenburg, SE-41296 Gothenburg, Sweden, EU 2Department of Physics, Chalmers University of Technology, SE-41296 Gothenburg, Sweden, EU 3Department of Mathematics and Program in Applied Mathematics, University of Arizona, Tucson, Arizona 85721, USA (Dated: August 1, 2018) The presence of a delay between sensing and reacting to a signal can determine the long-term behavior of autonomous agents whose motion is intrinsically noisy. In a previous work [M. Mijalkov, A. McDaniel, J. Wehr, and G. Volpe, Phys. Rev. X 6, 011008 (2016)], we have shown that sensorial delay can alter the drift and the position probability distribution of an autonomous agent whose speed depends on the illumination intensity it measures. Here, using theory, simulations, and experiments with a phototactic robot, we generalize this effect to an agent for which both speed and rotational diffusion depend on the illumination intensity and are subject to two independent sensorial delays. We show that both the drift and the probability distribution are influenced by the presence of these sensorial delays. In particular, the radial drift may have positive as well as negative sign, and the position probability distribution peaks in different regions depending on the delay. Furthermore, the presence of multiple sensorial delays permits us to explore the role of the interaction between them. PACS numbers: 05.60.-k, 05.40.Jc Keywords: autonomous agents, taxis, stochastic differential equations, sensorial delay I. INTRODUCTION Autonomous robots are increasingly being employed both in fundamental research and in technological set- tings [1]. One of the critical tasks in their development is to make them capable of complex autonomous behaviors in response to environmental cues, while keeping their hardware, sensorial inputs and software as simple as pos- sible [2, 3]. In fact, complex behaviors emerging from agents obeying simple rules have the advantage of being extremely robust and reliable [1, 4, 5]. Often, a source of inspiration are the behaviors of simple organisms like foraging insects [6] and chemotactic bacteria [7]. Usually, the robots are designed to react to real-time sensorial inputs from their surroundings and make de- cisions based on this information. In Nature, however, there are several examples of microscopic organisms and animals that compare current information about their surroundings with previous information, and adjust their behavior by making extrapolations. For example, chemo- tactic bacteria have been shown to adjust their motion by comparing the chemical concentration in their sur- roundings at different times [8, 9], and insects, fishes and humans extrapolate their positions forward in time when navigating in groups [10 -- 12]. These behaviors result in the introduction of a sensorial delay between the sensorial input perception and the ensuing behavioral response. We have recently explored the role played by this senso- rial delay both theoretically and experimentally [13, 14]. Using a phototactic robot whose speed depended on the measured light intensity, we demonstrated that introduc- ∗ These two authors contributed equally. ing a sensorial delay could make the robot either stay near or avoid the light source; furthermore, when multi- ple light-emitting robots interacted, we showed that this effect promoted either aggregation or segregation. The presence of negative sensorial delay, sometimes called "anticipation", has also been shown to greatly influence the dynamics of a system of interacting agents and the patterns that are formed [15], as well as to affect the clus- tering tendencies of agents in a two-dimensional variant of the Vicsek model [16]. Here, using theory, simulations, and experiments with phototactic robots, we generalize the effect we described in Ref. [13] to the case of an agent whose speed and rota- tional diffusion depend on the illumination intensity and are subject to two independent sensorial delays. Using a phototactic robot moving within an arena illuminated with a radial light intensity pattern, we investigate how the robot's behavior is affected by a delay when only its speed varies as a function of the intensity, only its rotational diffusion varies, or both quantities vary simul- taneously. We show that both its drift and its position probability distribution are influenced by the presence of these sensorial delays. In particular, the radial drift may have positive as well as negative sign, and the position probability distribution may peak in different regions de- pending on the delay. The presence of multiple sensorial delays permits us to explore the role of the interaction between them. II. MODEL The robot we employ can be modelled as an au- tonomous agent performing active Brownian motion [17]: it moves in the xy-plane while its orientation is subject 8 1 0 2 l u J 1 3 ] h p - o i b . s c i s y h p [ 1 v 5 6 7 1 1 . 7 0 8 1 : v i X r a 2 speed vmin and a maximum speed vmax, and decreases with higher light intensity, i.e.: v(I) = vmin + (vmax − vmin)e−I . (3) Furthermore, we now vary also the normalized rotational diffusion coefficient, so that it is bounded between mini- mum and maximum values Rmin and Rmax, and increases with higher light intensities, i.e.: R(I) = Rmax − (Rmax − Rmin)e−I . (4) We finally introduce the sensorial delays so that SDEs (2) become the stochastic differential delay equa- FIG. 1. Model and experimental setup. (a) The robot is at position (xt, yt) at time t and moves at a speed v with orienta- tion φt as indicated by the arrow. This orientation is subject to a noise so that the robot's characteristic reorentiation time is τ . (b) Picture of the robot in the arena illuminated with a light field generated by an infrared lamp. The robot is free to move in the region between the round object placed at the center of the arena and the black line on the outer edge of the arena. Depending on the scenario, the robot will either modify its speed (Fig. 2), its rotational diffusion (Fig. 3), or both simultaneously (Figs. 4 and 5) as functions of the light intensity it measures. tions (SDDEs): dxt dt dyt dt dφt dt = v(It−δv ) cos φt = v(It−δv ) sin φt (5) (cid:114) = 2R(It−δR) τ ηt where δv is the sensorial delay of the adjustment of the speed and δR is the sensorial delay of the adjustment of the rotational diffusion coefficient. A positive delay value corresponds to a delay in the time it takes to react to sensorial input, while a negative value corresponds to making a prediction of a future measured intensity [13 -- 15]. (1) III. THEORY to noise (Fig. 1a). Its behavior can therefore be modeled by the following stochastic differential equations (SDEs) [18]:  dxt dt dyt dt dφt dt = v cos φt = v sin φt (cid:114) 2 = ηt τ where (xt, yt) is the robot's position at time t, φt is its orientation, v is its speed, τ is its characteristic reorien- tation time (i.e., the time during which its orientation varies on average by one radiant), and ηt is a normally distributed white noise term with zero mean and unit intensity. Let us assume that the arena where the robot moves is illuminated by a light intensity I(x, y). If the robot can measure I and react to this measurement by adjusting its speed v and rotational diffusion R, the SDEs describing its motion become: dxt dt dyt dt dφt dt = v(It) cos φt = v(It) sin φt (2) (cid:114) = 2R(It) τ ηt where It = I(xt, yt). As in our previous work [13], we will assume that the robot speed is bounded by a minimum We theoretically study SDDEs (5) using multiscale analysis and derive expressions for the drift and steady- state position probability distribution of the robot. The multiscale analysis is a homogenization technique that is performed by taking to zero the characteristic time scales of the processes involved in determining the dynamics of the system, while keeping their ratios constant [19]. The detailed derivations are provided in Appendix A, while here we provide only an outline of the derivation and the key theoretical results. A. Outline of the derivation We start by rewriting SDDEs (5) in a more convenient form for the theoretical analysis, introducing, in particu- lar, a parameter  that will be taken to zero in the mul- tiscale analysis [19]. We note that the speed of the robot is a function of its position, i.e., v(x, y) = v(I(x, y)), and that the robot changes the direction of its velocity accord- ing to a random process, at a rate which is also a function If the robot reacts to the environment with a delay δv = c2, the speed at time t is proportional to v (xt−c2, yt−c2) of the position, i.e., σ(x, y) = (cid:112)2R(I(x, y))/τ . 01002003000100200300ab (the value of the function v evaluated at the position of the particle at an earlier moment of time, if c > 0, or at a later moment, if c < 0). Likewise, the rate of the robot random rotation is proportional to σ (xt−k2, yt−k2), with a delay δR = k2. The parameters c and k are constants, positive or negative, and, in general, different from one another. Thus, we can rewrite SDDEs (5) as a set of SDDEs with a small parameter :  dxt = dyt = dφt = 1  1  1  v (xt−c2, yt−c2) cos φt dt v (xt−c2, yt−c2 ) sin φt dt (6) σ (xt−k2 , yt−k2 ) dWt where Wt, t ≥ 0, denotes a Wiener process.1 Since the factor of 1  in the equation for φt makes the changes of direction occur very rapidly for  small, we scale the speed v in the first two equations in the same way to obtain a nontrivial limiting dynamics for the position of the robot. We remark that the SDDEs (6) becomes SDDEs (5) for  = 1. We study the limit of SDDEs (6) for  → 0, which is equivalent to accelerating the microscopic dynamics (speed, rotation, delays) of the system while keeping its macroscopic properties (drift, probability distribution) fixed. We first linearize x and y as functions of time, and then v and σ as functions of x and y, to approximate the SDDEs (6) by a system of SDEs without delays. We then consider the corresponding backward Kolmogorov equation for the probability density ρ, write the function ρ as a formal series in powers of , i.e., ρ = ρ0 + ρ1 + 2ρ2 + . . . , (7) and use the multiscale expansion method to derive the backward Kolgomorov equation for the limiting density ρ0: ∂ρ0 ∂t (cid:26) (cid:26) (cid:20) v = − 1 2R δv σ2 v τ 2 v2 2R δR σ2 σ τ + (cid:26) ∂ + v ∂x σ2 ∂ρ0 ∂x ∂ρ0 ∂x (cid:21) σx ∂ ∂x vx + vy + σy (cid:27) (cid:27) (cid:20) v ∂ρ0 ∂y ∂ρ0 ∂y ρ0 + ∂ ∂y ∂ ∂y σ2 (cid:18) σ2 ∇· 2R (cid:19) +∇·(cid:16) v (cid:17) σ2∇ (vρ0) , by the processes xt and yt:  dxt = dyt = (cid:20) (cid:20) − 1 2 √ + 2 − 1 2 √ + 2 δv τ v σ δv τ v σ δR τ 2R σ2 v2 σ σx + v ∂ ∂x δR τ 2R σ2 v2 σ σy + v ∂ ∂y 2R σ2 vvx + dW (1) t 2R σ2 vvy + dW (2) t 3 dt dt (cid:16) v (cid:16) v σ2 (cid:17)(cid:21) (cid:17)(cid:21) σ2 (9) where W (1) and W (2) are independent Wiener processes. From SDEs (9), we obtain the associated forward Kol- mogorov (or Fokker-Planck) equation ∂tρ0 = 1 2 δv τ 2R σ2 ∇·(ρ0v∇v)− δR τ ρ0 ∇σ v2 σ (10) from which the stationary probability density ρ0 can be found by solving for ∂tρ0 = 0. B. Key results in circular geometry Given the circular geometry of our experiment (see Section IV), we can assume v and σ in SDEs (9) and in Eq. (10) to be rotationally invariant. We can there- fore study these equations in polar coordinates focusing specifically on the radial coordinate r =(cid:112)x2 + y2.2 We therefore obtain the following homogenized SDE for the radial coordinate: drt = d(r) dt + s(r) d Wt, (11) where Wt is a Wiener process, the radial drift coefficient is d(r) = − 1 2 δv τ 2R σ2 vvr + δR τ 2R σ2 v2σr σ + v and the noise coefficient is s(r) = √ 2 v(r) σ(r) . (cid:16) v (cid:17) σ2 r + 1 r v2 σ2 (12) (13) (cid:21)(cid:27) ρ0 . (8) The steady-state radial probability distribution is (cid:26)(cid:90) 2b(r) s2(r) (cid:27) ρ0(r) = B s2(r) exp dr , (14) From this equation, we derive the limiting SDEs satisfied where B has to be adjusted to make the integral of ρ equal 1. 1 The stochastic differential in the third equation can also be writ- ten as dWt = ηt dt, where ηt, t ≥ 0, is a unit white noise process. 2 The results for the azimuthal coordinate are trivial: because of rotation symmetry, the azimuthal drift must be null and the azimuthal position probability distribution must be uniform. IV. ROBOT EXPERIMENT The experimental setup is shown in Fig. 1b. We use an Elisa-3 [20] robot, which is an autonomous robot with a circular shape that measure 50 mm in diameter and 30 mm in height. The robot moves at a maximum speed of 60 cm s−1 thanks to two wheels on either side powered by direct-current (DC) motors. It is equipped with eight infrared (IR) sensors that measure ambient light placed along the perimeter of the robot at equal intervals of 45 degrees to create a detection field of 360 degrees. Fur- thermore, the robot features proximity sensors that per- mit it to detect the presence of objects at a distance of 6 cm and four ground sensors that permit it to detect the presence of a black border on the ground. We have programmed the robot using Aseba studio [21]. The robot can perform Brownian motion through a cycle of two phases: a "forward phase" when the robot moves forward at constant speed along a straight line for 0.1 s; and a "rotation phase" when the robot changes its direction by a random angle for 0.1 s. This cycle is repeated to emulate a Brownian motion. We delimit a region where the robot can move freely by placing a circular object at the center of the arena and a black tape along its outer edge, as shown in Fig. 1b. The robot uses the proximity sensors to detect the circular object and the ground sensors to detect the black tape, and it avoids them by changing its direction away from them, i.e. until it does not detect their presence any more. We generate a radially decaying light intensity field by placing a 150-W IR lamp above the arena where the robot moves. The robot measures the local value of this light intensity using the IR sensors and adapts its behavior accordingly. Following the approach in our previous work [13], we estimate the values of It−δv and It−δR by an expansion to the first order, i.e., I(t− δv) = I(t)− δvI(cid:48)(t) and I(t − δR) = I(t) − δRI(cid:48)(t), respectively. Practically, the robot stores the value of the intensity in the previous and current motion cycles, and uses them to approximate the intensity derivative. During the experiments, the robot's positions are recorded with a videocamera at 32 fps and tracked us- ing standard digital video microscopy algorithms. Each experiment runs for 60 minutes. From the acquired trajectories, we estimate the radial probability distribution of the robot's position and its radial drift. The radial probability distribution is the probability of finding a robot at a certain radial distance from the center of the arena and is directly measured from the histogram of the robot's positions. The radial drift shows how the robot moves on average relative to the center of the arena depending on its location and is measured using the following equation [13]: d(r) = 1 ∆t (cid:104)rn+1 − rnrn ∼= r(cid:105), (15) where rn is the series of robot's positions and ∆t is the 4 time step. If the radial drift is positive, the robot on average moves away from the center of the arena, whereas a negative drift means that it moves on average towards the center. V. RESULTS We consider three scenarios. First, we vary only the speed as a function of light intensity (as in our previous work [13]). Second, we vary the rotational diffusion co- efficient. Third, we vary both quantities simultaneously so that the presence of multiple sensorial delays permits us to explore how they interact. In all cases, we present the theoretical, simulation and experimental results. The simulations are realized by a finite-difference algorithm [17] that implements SDDEs (5) using the experimental parameters.3 A. Speed dependent on the light intensity We set the speed to vary between the vmax = 25.7 mm s−1 and vmin = 4.3 mm s−1 according to Eq. (3) (Fig. 2a), while the rotational diffusion is kept constant at DR = τ−1 = 0.29 rad2s−1 (Fig. 2b). This case is equivalent to that we had previously studied [13]. The qualitative behavior of the robot can be seen from its trajectories in the presence of different sensorial de- lays. In the absence of any delay (Fig. 2d), the robot has a slight preference to spend time in the regions with low speed (corresponding to high light intensity). This tendency is accentuated when a positive sensorial delay is introduced (δv = +10τ , Fig. 2c), while it can be re- versed by introducing a sufficiently large negative delay (δv = −10τ , Fig. 2e). These qualitative observations can be made more pre- cise by measuring the radial probability distribution ρ0(r) (Figs. 2f-h) and the radial drift d(r) (Figs. 2i-k) of the robot in each case. The theoretical results (solid lines) agree well with the simulations (dashed lines) and experiments (symbols). As we qualitatively discussed above (Figs. 2c-e), the sensorial delay δv influences the robot probability distribution ρ0(r) so that ρ0(r) peaks in the regions with higher light intensity and lower speed for δv = +10τ (Fig. 2f), and in the regions with lower light intensity and higher speed for δv = −10τ (Fig. 2h). The radial drift d(r) is also influenced by δv so that d(r) is mostly negative when δv = +10τ pulling the robot to- wards the center of the arena (Fig. 2i), and it is positive 3 As in the experiments, we introduce the sensorial delays by esti- mating the values of It−δv and It−δR by an expansion to the first order, i.e., I(t−δv) = I(t)−δvI(cid:48)(t) and I(t−δR) = I(t)−δRI(cid:48)(t), respectively. 5 FIG. 2. Robot behavior with sensorial delay in the speed. (a) Speed v(r) and (b) rotational diffusion coefficient DR(r) as a function of radial position. (c-e) 60-minute-long trajectories of the robot within the arena for positive, zero and negative delays. (f-h) Radial probability distributions ρ0(r)/r and (i-k) radial drift d(r) of the robot for positive, zero and negative delays; the symbols represent experimental data with standard deviation, the dashed lines represent simulations, and the solid lines represent the theory (Eqs. (12) and (14)). when δv = −10τ pushing the robot towards the edge of the arena (Fig. 2k).4 We observe that there are significant deviations be- tween the theoretical ρ0(r) and d(r) and those obtained from experiments and simulations, especially towards the edges of the arena. These deviations emerge because the experiments and simulations implement SDDEs (5), cor- responding to  = 1 in SDDEs (6), while the theory is strictly valid for  → 0. This is discussed in more detail in Section V D. 4 Note that in this case (i.e. constant rotational diffusion) the critical value where the sign change of d(r) occurs is δv = − 2 σ , as we have shown in Ref. [13]. B. Rotational diffusion dependent on the light intensity In this second case, we set the rotational diffusion to vary between DR,max = 1.4 rad2s−1 and DR,min = 0.014 rad2s−1 (Fig. 3b) according to Eq. (4), while keep- ing v = 17.1 mm s−1 (Fig. 3a). Figs. 3c-e show the trajectories of the robot for various sensorial delays. For positive delay (δR = +10τ , Fig. 3c), the robot spends most of its time close to the center of the arena, where the light intensity and the rotational diffu- sion take larger values. For zero delay (δR = 0, Fig. 3d), in full agreement with Eq. (A36), the space explored by the robot does not seem to be influenced by the light in- tensity and rotational diffusion values. For negative delay (δR = −10τ , Fig. 3e), the robot spends most of its time in the region close to the edge of the arena, where the light intensity and the rotational diffusion take smaller 111722a020040000.51bcde00.20.40.6fTheorySimulationExperimentgh100300500-10010i100300500j100300500k 6 FIG. 3. Robot behavior with sensorial delay in the rotational diffusion. (a) Speed v(r) and (b) rotational diffusion coefficient DR(r) as a function of radial position. (c-e) 60-minute-long trajectories of the robot within the arena for positive, zero and negative delays. (f-h) Radial probability distributions ρ0(r)/r and (i-k) radial drift d(r) of the robot for positive, zero and negative delays; the symbols represent experimental data with standard deviation, the dashed lines represent simulations, and the solid lines represent the theory (Eqs. (12) and (14)). values. The radial probability distribution ρ0(r) (Figs. 3f-h) and the radial drift d(r) (Figs. 3i-k) confirm these qual- itative observations. In particular, we observe that the theoretical ρ0(r) when δR = 0 corresponds to a uniform distribution (solid line in Fig. 3g), and the correspond- ing d(r) is almost zero (solid line in Fig. 3j). For δR > 0, ρ0(r) is peaked towards the high intensity and rotational diffusion regions near the center of the arena (Fig. 3f), and d(r) assumes mostly negative values, pulling the robot towards the arena center (Fig. 3i). For δR > 0, the reverse is true: ρ0(r) is peaked towards the low in- tensity and rotational diffusion regions near the edge of the arena (Fig. 3f), and d(r) assumes positive values, pushing the robot towards the arena edge (Fig. 3i).5 Also in this case there are some discrepancies between the theory (gray lines) and the simulations (dashed lines) and experiments (symbols), which can be explained by the fact that simulations and experiments are not realized at the limit for  → 0 (see Section V D). C. Both speed and rotational diffusion dependent on the light intensity It is also interesting to consider the hybrid cases when both the speed and the rotational diffusion depend on the light intensity. In Fig. 4, we consider the case where both sensorial delays have the same sign, and reinforce 5 Note that in this case (i.e. constant speed) the critical value where the sign change of d(r) occurs is δR = 0. 111722a020040000.51bcde00.20.40.6fTheorySimulationExperimentgh100300500-20-1001020i100300500j100300500k 7 FIG. 4. Robot behavior with equal sensorial delays in both the speed and the rotational diffusion. (a) Speed v(r) and (b) rotational diffusion coefficient DR(r) as a function of radial position. (c-e) 60-minute-long trajectories of the robot within the arena for positive, zero and negative delays. (f-h) Radial probability distributions ρ0(r)/r and (i-k) radial drift d(r) of the robot for positive, zero and negative delays; the symbols represent experimental data with standard deviation, the dashed lines represent simulations, and the solid lines represent the theory (Eqs. (12) and (14)). each other. Figs. 4a and 4b show v(r) and DR(r), re- spectively. Some samples of the resulting trajectories are shown in Figs. 4c-e. The two sensorial delays reinforce each other and produce a more pronounced effect on the way the particle explores the space: when δv = +10τ and δR = +10τ , the robot is attracted towards the high- light-intensity regions at the center of the arena where the speed is low and the rotational diffusion is high (Fig. 4c); when δv = −10τ and δR = −10τ , the robot moves to- wards the low-light-intensity regions near the edges of the arena where the speed is high and the rotational diffusion is low (Fig. 4e). This enhancement of the robot motion is further confirmed by the changes in the corresponding ρ0(r) (Figs. 4f-h) and d(r) (Figs. 4i-k). tively. Some sample trajectories are shown in Figs. 5c and 5d: when δv = +10τ and δR = −10τ , the robot is attracted towards the regions with low-light-intensity regions where the speed is high and the rotational diffu- sion is low (Fig. 5c); when δv = −10τ and δR = +10τ , the robot is attracted towards the high-light-intensity re- gions where the speed is low and the rotational diffusion is high (Fig. 5d). These results are supported by ρ0(r) (Figs. 5e-f) and D(r) (Fig. 5g-h). D. Differences between theory and simulations/experiments In Fig. 5, we consider the case when the two senso- rial delays have opposite signs, and compete with each other. Figs. 5a and 5b show v(r) and DR(r), respec- In all data presented in Figs. 2, 3, 4, and 5, we ob- tain a very good agreement between the experiments and simulations, while there are certain discrepancies when 111722a020040000.51bcde00.20.40.6fTheorySimulationExperimentgh100300500-30-1501530i100300500j100300500k 8 FIG. 5. Robot behavior with opposite sensorial delays in the speed and the rotational diffusion. (a) Speed v(r) and (b) rotational diffusion coefficient DR(r) as a function of radial position. (c-d) 60-minute-long trajectories of the robot within the arena. (e-f) Radial probability distributions ρ0(r)/r and (g-h) radial drift d(r) of the robot; the symbols represent experimental data with standard deviation, the dashed lines represent simulations, and the solid lines represent the theory (Eqs. (12) and (14)). it comes to the theory, particularly for the cases of the drift with negative delays. This can be explained taking into consideration that the theory assumes taking the value  → 0 in SDDEs (6) (while keeping δv/τ and δR/τ constant), while the simulations and experiments are per- formed at the finite value of  = 1. We tested this hypoth- esis by running simulations where  was taken towards zero while keeping the ratio between the time scales of the system the same as in the experiments. In Fig. 6, the results can be observed for a simulated robot whose speed varies as a function of the intensity under the influence of a negative delay of δv = −10τ . The radial probability dis- tribution of the robot can be seen to converge towards the theoretical distribution when  → 0 (Figs. 6a-c). An even more significant change can be observed for the robot's radial drift: while for  = 1 (Fig. 6d) there is a sig- nificant difference between the simulated and theoretical radial drifts, this difference is significantly reduced when  = 0.5 (Fig. 6e) and, even more, when  = 0.1 (Fig. 6f). Note also that in all cases there is a sharp drop in the simulated robot radial drift near the boundary, which is due to the fact that the theory does not account for the boundary's presence. VI. CONCLUSIONS We have explored the role that sensorial delays play in determining the motion of an autonomous robot. Ex- tending our previous work [13], we have considered a phototactic robot whose speed and rotational diffusion depend on the local value of the light intensity. We have shown that the introduction of sensorial delays leads to an alteration of both the position probability distribu- tion and the drift of the robot. These results can be used to engineer the motion of autonomous agents using sen- 161820a020040000.51bcd00.20.40.6eTheorySimulationExperimentf100300500-20-1001020g100300500h 9 FIG. 6. Convergence of the simulation results to the theoretical predictions for a simulated agent with sensorial delay in the speed. The simulated (a-c) radial probability distribution ρ0(r)/r and (d-f) radial drift d(r) converge towards the theoretical prediction as  → 0 (while keeping the δv/τ and δR/τ constant). sorial delays as well as to explain how multiple sensorial delays can interplay to obtain the desired behavior of a system. ACKNOWLEDGMENTS We thank Lovisa Hagstom, Erik Holmberg, Eliza Nord´en, Teodor Norrestad, Martin Selin and Lisa Sjoblom for performing an early version of the simula- tions presented in this work during their Bachelor Thesis, as well as Mite Mijalkov and Gilles Caprari for useful dis- cussions. This work was partially supported by the ERC Starting Grant ComplexSwimmers (Grant No. 677511). JW's work was partially supported by the NSF grant DMS-1615045. Appendix A: Mathematical Derivation Starting from SDDEs (6), we approximate them with a system of SDEs without delays by linearizing x and y as functions of time, and then v and σ as functions of x and y. As a result, we obtain: v (xt−c2 , yt−c2 ) ≈ v (xt, yt)−vx (xt, yt) c2 xt−vy (xt, yt) c2 yt (A1) with vx and vy denoting the partial derivatives of v, and dots denoting time derivatives. Substituting this expres- sion into the first two equations of SDDEs (6), we obtain  xt = yt = yt =  xt =  xt = yt = 1  1  1  1  approximate versions of these equations: v cos φt − cvx cos φt xt − cvy cos φt yt v sin φt − cvx sin φt xt − cvy sin φt yt 1  1  (A2) From this point on, v, vx and vy are always evaluated at (xt, yt) and we omit their arguments from the notation. Eqs. (A2) constitute a system of linear equations for xt and yt, whose solution is v cos φt [1 + c (vx cos φt + vy sin φt)] −1 v sin φt [1 + c (vx cos φt + vy sin φt)] −1 (A3) For small , we can approximate further, obtaining the first two equations of the system we will study: v cos φt [1 − c (vx cos φt + vy sin φt)] v sin φt [1 − c (vx cos φt + vy sin φt)] (A4) To obtain the third equation, we start from a similar approximation of the function σ: σ (xt−k2, yt−k2) ≈ σ (xt, yt)−σx (xt, yt) k2 xt−σy (xt, yt) k2 yt. (A5) 00.20.40.6aTheorySimulationbc0200400-10010d0200400e0200400f v cos φt v sin φt (A6) L0 = We further approximate the expression on the right-hand side, replacing xt and yt by their leading order terms6 from Eqs. (A6):  xt ≈ 1 (cid:20) 1  yt ≈ 1  dφt =  (cid:21) to obtain the third equation of the approximate system: σ (xt, yt) − kσxv cos φt − kσyv sin φt dWt. (A7) In order to study the limiting behavior of the pro- cess (xt, yt), we introduce the associated (backward) Kol- mogorov operator.7 In our case, (cid:21)2 (cid:20) 1 (cid:18) 1 (cid:18) 1    + + L = 1 2 σ − kσxv cos φ − kσyv sin φ ∂2 φφ v cos φ − cvvx cos2 φ − cvvy cos φ sin φ v sin φ − cvvx sin φ cos φ − cvvy sin2 φ (cid:19) (cid:19) ∂y. (A8) Considering the corresponding backward Kolmogorov equation for a function p(t, x, y, φ), ∂tρ = Lρ, (A9) we have L = −2L−2 + −1L−1 + L0, (A10) 6 Including higher order terms in the approximations for xt and yt, substituted into Eqs. (A5), would give rise to terms of or- der  in the equation for φt (Eq. (A7)). As can be seen from the asymptotic analysis that follows, this would not change the equation obtained in the  → 0 limit. 7 The general rule is the following: consider a system of SDE dxi t = bi(xt) dt + σi α(xt) dW α t l(cid:88) α=1 where i = 1, . . . k and W 1, . . . W l are independent Wiener pro- cesses. The generator is then the differential operator k(cid:88) i,j=1 k(cid:88) i=1 L = 1 2 aij (x) ∂2 ∂xi∂xj + bi(x) ∂ ∂xi aij =(cid:80) where the aij are matrix elements of the matrix a = σσT , i.e. α σi ασj α. Consult [18] for more details. where 10 1 2 σ2∂2 φφ L−2 = L−1 = (−kσσxv cos φ − kσσyv sin φ) ∂2 φφ k2σ2 xv2 cos2 φ + k2σxσyv2 cos φ sin φ + v cos φ∂x + v sin φ∂y (cid:18) 1 −(cid:0)cvvx cos2 φ + cvvy cos φ sin φ(cid:1) ∂x −(cid:0)cvvx sin φ cos φ + cvvy sin2 φ(cid:1) ∂y yv2 sin2 φ (cid:19) 2 1 2 k2σ2 ∂2 φφ + We write the function ρ as a formal series in powers of : ρ = ρ0 + ρ1 + 2ρ2 + . . . , (A12) (A11) substitute it into Eq. (A9), and equate coefficients of the same powers of  on both sides of the resulting equation. The goal is to obtain a differential equation for ρ0, which (in view of Eq. (A12)) is the limit of ρ as  → 0. In order −2, we obtain ∂x L−2ρ0 = 0, (A13) which implies ∂2 φφρ0 = 0. While the general solution of this equation is an affine function of φ, that is, has a form a(x, y) + b(x, y)φ, we choose a solution ρ0(x, y) which does not depend on φ, since we expect that the limiting equation does not involve the fast variable φ. In order −1, we have L−2ρ1 = −L−1ρ0, which implies the equation φφρ1 = − 2v ∂2 σ2 ∂xρ0 cos φ − 2v σ2 ∂yρ0 sin φ whose solution, periodic in φ, is ρ1 = 2v σ2 ∂xρ0 cos φ + 2v σ2 ∂yρ0 sin φ. In order 0, we obtain (A14) (A15) (A16) ∂tρ0 = L−2ρ2 + L−1ρ1 + L0ρ0. (A17) This can be rewritten as ∂tρ0 − L−1ρ1 − L0ρ0 = L−2ρ2, which implies that the function ∂tρ0 − L−1ρ1 − L0ρ0 be- longs to the range of the operator L−2, and is thus or- thogonal to the null space of the adjoint operator L∗ −2.8 L−2 is considered here as an operator in the variable φ. 8 This is a general fact about linear operators on Hilbert spaces. A discussion in the present context can be found e.g. in [19]. (cid:90) π −π ∂tρ0 = 1 2π Since σ does not depend on φ, L∗ −2 = L−2 and the null space of this operator is spanned by the constant function 1. The orthogonality relation becomes (L0ρ0 + L−1ρ1) dφ. (A18) Substituting L0ρ0 = −(cid:0)cvvx cos2 φ∂xρ0 + cvvy sin2 φ∂yρ0 (cid:1) (A19) and L−1ρ1 = 2kv2 σ (cid:18) 2v (cid:18) 2v (σx cos φ + σy sin φ) (∂xρ0 cos φ + ∂yρ0 sin φ) + v cos φ ∂x + v sin φ ∂y σ2 ∂xρ0 cos φ + σ2 ∂xρ0 cos φ + 2v σ2 ∂yρ0 sin φ 2v σ2 ∂yρ0 sin φ (cid:19) (cid:19) and using the trigonometric integrals cos2 φ dφ = 1 2π sin2 φ dφ = 1 2 (A21) (cid:90) π −π (cid:90) π −π 1 2π and cos φ sin φ dφ = 0, (A22) (cid:90) π −π 1 2π we obtain ∂tp0 = Lρ0 = − 1 2 + k + v∂x or, in vector notation, ∂tρ0 = − 1 2 cv∇v · ∇ρ0 + k cvvx∂xρ0 − 1 2 v2 σ σx∂xρ0 + k (cid:16) v (cid:17) σ2 ∂xρ0 v2 σ σy∂yρ0 (cid:16) v + v∂y σ2 ∂yρ0 (cid:17) (A23) ∇σ · ∇ρ0 + v∇ ·(cid:16) v σ2∇ρ0 (cid:17) . v2 σ (A24) This is the limiting backward Kolmogorov equation for a function ρ0 of the variables x and y, from which we obtain the system of SDEs, satisfied by the processes xt (cid:20) (cid:20) and yt: dyt = dxt = − 1 2 − 1 2 cvvx + k cvvy + k v2 σ v2 σ σx + v∂x σy + v∂y (cid:16) v (cid:16) v σ2 (cid:17)(cid:21) (cid:17)(cid:21) σ2 11 where W (1) and W (2) are independent Wiener processes.9 Note that in the case when σ is identically equal to 1, we obtain the system studied previously in Ref. [13]. Passing to formal adjoints, we obtain the associated forward Kolmogorov (i.e., Fokker-Planck) equation: ∂tρ0 = L∗ρ0 = 1 2 (cid:19) (cid:17) 1 c∂x (vvxρ0) + 2 − k∂x − ∂x (cid:18) v2 (cid:16) (cid:16) v (cid:19) (cid:18) v2 v∂x σ2 σ σxρ0 c∂y (vvyρ0) − k∂y (cid:18) v2 (cid:17) − ∂y (cid:16) (cid:18) v2 σ ρ0 v∂y (cid:19) σ2 ρ0 + ∂2 yy σ2 ρ0 , + ∂2 xx (cid:19) (cid:16) v σ2 σyρ0 (cid:17) (cid:17) ρ0 which, in vector form, becomes , ∂tρ0 = 1 2 c∇·(ρ0v∇v)−k∇· ρ0 ∇σ v2 σ (cid:18) (A26) (cid:19) +∇·(cid:16) v (cid:17) σ2∇ (vρ0) . (A20) (A27) If the system possesses a stationary probability density ρ0, then ρ0 has to satisfy the stationary Fokker-Planck equation L∗ρ0 = 0 . (A28) We are now going to consider some special cases, where the solutions of the stationary Fokker-Planck equation actually satisfy a stronger condition. Ref. [22] explains in detail that in these situations we actually obtain an equi- librium distribution, i.e., a stationary distribution satis- fying the detailed balance condition. cvvy∂yρ0 a. Constant rotational diffusion Suppose σ is constant, which correspond to the special case we studied in Ref. [13]. In this case the stationary Fokker-Planck equation becomes c∇ · (ρ0v∇v) + 1 2 1 σ2∇ · (v∇ (vρ0)) = 0. (A29) We search for a solution of cρ0v∇v + 1 2 1 σ2 v∇ (vρ0) = 0, which can be rewritten as ∇ (vρ0) vρ0 = − σ2 2 c ∇v v (A30) (A31) dt + dt + √ √ 2 v σ dW (1) t 2 dW (2) v σ (A25) t 9 This step is a reversal of the previous operation by which we obtained a Kolmogorov equation from an SDE system. Having passed to the limit at the level of Kolmogorov equations, we revert back to the corresponding SDEs. and integrated to yield ρ0 = Bv −(cid:16) 1+ σ2c 2 (cid:17) . (A32) In a bounded domain, a positive value of B can always be chosen, so as to make ρ0 a probability distribution. For c > − 2 σ2 , the points (x, y) with smaller values of v are preferred by this distribution; for c < − 2 σ2 the tendency is reversed. b. Constant speed Suppose v is constant. In this case the stationary Fokker-Planck equation becomes (cid:18) − k∇ · ρ0 ∇σ v2 σ (cid:19) + ∇ ·(cid:16) v (cid:17)∇ (vρ0) = 0. σ2 A function ρ0 will satisfies this equation if it satisfies the equation − kv2 ρ0 σ ∇σ + v σ2∇ (vρ0) = 0. For this, it is enough to find a solution of ∇ (vρ0) vρ0 = kσ∇σ, i.e., ∇ log (vρ0) = which has a general solution ρ0 = B exp 1 2 k∇(cid:0)σ2(cid:1) , (cid:19) (cid:18) 1 kσ2 . 2 (A33) (A34) (A35) (A36) (A37) This can be normalized to become a probability density if the point (x, y) is restricted to a bounded domain. For k > 0, it shows that the particle is more likely to be found in the region where its rate of rotation is bigger. For k < 0 it has the opposite tendency. c. Speed proportional to the rotational diffusion Yet another case in which the stationary Fokker-Planck equation can be integrated explicitly is the case when v/σ2 is a constant, i.e., when the speed is proportional to the rotational diffusion. The calculation is straight- forward. Suppose both v and σ are functions of r =(cid:112)x2 + y2, d. Radial coordinates which is the special case of the experiment we performed. We are going to find an SDE satisfied by rt =(cid:112)x2 the Ito formula for the function r =(cid:112)x2 + y2, t + y2 t and use it to derive the stationary distribution of the particle's distance from the origin. To this end, we use 12 drt = xt rt dxt+ yt rt dyt+ 1 2 t − x2 r2 r3 t t (dxt)2+ 1 2 t − y2 r2 r3 t t (dyt))2, (A38) and substitute the expressions for dxt and dyt from Eqs. (A25). Noting that for a function f (x, y) we have r fx + y fr = fxxr + fyyr = x r fy, and using the fact that xt is a differential of a Wiener process, rt which we will denote by Wt, we obtain t + yt rt dW (1) dW (2) t (cid:16) v (cid:17) σ2 r (cid:19) + 1 r v2 σ2 √ dt+ 2 v σ d Wt. (cid:18) drt = − 1 2 cvvr + k v2σr σ + v (A39) Denoting the drift and the noise coefficients in the above SDE by b(r) and s(r) respectively, we have the standard formula for the density ρ0 of the stationary distribution of r: (cid:18)(cid:90) r (cid:19) ρ0(r) = B s2(r) exp 2b(u) s(u)2 r0 du. (A40) Here, r0 > 0 is the minimal distance of the particle from the origin, allowed by the experimental restrictions [22] and B is the constant, normalizing the integral of g to 1. The integrand 2b s2 can be written explicitly as 2b s2 = − 1 2 c σ2vr v + kσσr + σ2 v + 1 r . (A41) (cid:16) v (cid:17) σ2 r Here is a short derivation of Eq. (A40): according to the general rule, the generator of the process rt is given by L = 1 2 s(r)2 d2 dr2 + b(r) d dr ; (A42) the adjoint generator is thus acting on functions of r ac- cording to the formula (L∗f )(r) = 1 2 d2 dr2 (cid:0)s(r)2f (r)(cid:1) − d dr (b(r)f (r)) ; (A43) to find the stationary density we solve the equation L∗ρ0 = 0, searching for a solution, satisfying the (stronger) equation (cid:0)s(r)2ρ0(r)(cid:1) − (b(r)ρ0(r)) = 0; (A44) 1 2 d dr to solve this first-order ODE, we substitute g = s2ρ0 and obtain which can be solved by separation of variables, i.e., (A45) − b 1 2 dg dr s2 g = 0, (cid:18)(cid:90) 2b(r) s2(r) (cid:19) g = B exp dr (A46) where B is a constant; from this formula, we obtain ρ0(r) = B s2(r) exp dr (A47) (cid:18)(cid:90) 2b(r) s2(r) (cid:19) where B has to be adjusted to make the integral of ρ0 equal 1. 13 [1] F. Schweitzer. Brownian agents and active particles. 205:3321, 2002. Springer, 2003. [2] M. Rubenstein, A. Cornejo, and R. Nagpal. Pro- grammable self-assembly in a thousand-robot swarm. Science, 345:795, 2014. [3] J. Werfel, K. Petersen, and R. Nagpal. Designing col- lective behavior in a termite-inspired robot construction team. Science, 343:754, 2014. [4] J. Palacci, S. Sacanna, A. P. Steinberg, D. J. Pine, and P. M. Chaikin. Living crystals of light-activated colloidal surfers. Science, 339:936, 2013. [5] O. Chepizhko and F. Peruani. Diffusion, subdiffusion, and trapping of active particles in heterogeneous media. Phys. Rev. Lett., 111:160604, 2013. [6] E. Bonabeau, M. Dorigo, and G. Theraulaz. Inspiration for optimization from social insect behaviour. Nature, 406:39, 2000. [13] M. Mijalkov, A. McDaniel, J. Wehr, and G. Volpe. En- gineering sensorial delay to control phototaxis and emer- gent collective behaviors. Phys. Rev. X, 6:011008, 2016. [14] G. Volpe and J. Wehr. Effective drifts in dynamical systems with multiplicative noise: a review of recent progress. Rep. Prog. Phys., 79:053901, 2016. [15] P. Gerlee, K. Tunstrøm, T. Lundh, and B. Wennberg. Impact of anticipation in dynamical systems. Phys. Rev. E, 96:062413, 2017. [16] R. Piwowarczyk, M. Selin, T. Ihle, and G. Volpe. Influ- ence of sensorial delay on clustering and swarming. arXiv, page 1803.06026, 2018. [17] G. Volpe, S. Gigan, and G. Volpe. Simulation of the ac- tive Brownian motion of a microswimmer. Am. J. Phys., 82:659, 2014. [18] B. Øksendal. Stochastic differential equations. Springer [7] H. C. Berg. E. coli in Motion. Springer Science & Busi- Science & Business Media, 2003. ness Media, 2008. [8] R. M. Macnab and D. E. Koshland. The gradient-sensing mechanism in bacterial chemotaxis. Proc. Natl. Acad. Sci. USA, 69:2509, 1972. [9] J. E. Segall, S. M. Block, and H. C. Berg. Temporal comparisons in bacterial chemotaxis. Proc. Natl. Acad. Sci. USA, 83:8987, 1986. [10] T. S. Collett and M. F. Land. How hoverflies compute interception courses. J. Comp. Physiol., 125:191, 1978. [11] R. Nijhawan. Motion extrapolation in catching. Nature, 370:256, 1994. [12] S. Rossel, J. Corlija, and S. Schuster. Predicting three- dimensional target motion: How archer fish determine where to catch their dislodged prey. J. Exp. Biol., [19] G. Pavliotis and A. Stuart. Multiscale methods: Aver- aging and homogenization. Springer Science & Business Media, 2008. [20] Elisa-3 - gctronic wiki. http://www.gctronic.com/doc/ index.php/Elisa-3. Accessed: 06-04-2018. [21] Aseba studio - thymio & aseba. https://www.thymio. org/en:asebastudio. Accessed: 27-04-2018. [22] J. Birrel and J. Wehr. A homogenization theorem for Langevin systems with an application to Hamiltonian dynamics. arXiv, page 1707.02884, 2017. To appear in Charles M. Newman Festschrift (V. Sidoravicius, ed.), Springer (2018)".
1201.0705
2
1201
2012-01-04T17:43:46
Diversity of Dynamics and Morphologies of Invasive Solid Tumors
[ "physics.bio-ph" ]
Complex tumor-host interactions can significantly affect the growth dynamics and morphologies of progressing neoplasms. The growth of a confined solid tumor induces mechanical pressure and deformation of the surrounding microenvironment, which in turn influences tumor growth. In this paper, we generalize a recently developed cellular automaton model for invasive tumor growth in heterogeneous microenvironments [Y. Jiao and S. Torquato, PLoS Comput. Biol. 7, e1002314 (2011)] by incorporating the effects of pressure. Specifically, we explicitly model the pressure exerted on the growing tumor due to the deformation of the microenvironment and its effect on the local tumor-host interface instability. Both noninvasive-proliferative growth and invasive growth with individual cells that detach themselves from the primary tumor and migrate into the surrounding microenvironment are investigated. We find that while noninvasive tumors growing in "soft" homogeneous microenvironments develop almost isotropic shapes, both high pressure and host heterogeneity can strongly enhance malignant behavior, leading to finger-like protrusions of the tumor surface. Moreover, we show that individual invasive cells of an invasive tumor degrade the local extracellular matrix at the tumor-host interface, which diminishes the fingering growth of the primary tumor. The implications of our results for cancer diagnosis, prognosis and therapy are discussed.
physics.bio-ph
physics
Diversity of Dynamics and Morphologies of Invasive Solid Tumors Yang Jiao∗ Physical Science in Oncology Center, Princeton University, Princeton New Jersey 08544, USA Salvatore Torquato† Department of Chemistry, Princeton University, Princeton New Jersey 08544, USA Department of Physics, Princeton University, Princeton New Jersey 08544, USA Physical Science in Oncology Center, Princeton University, Princeton New Jersey 08544, USA Princeton Center for Theoretical Science, Princeton University, Princeton New Jersey 08544, USA and Program in Applied and Computational Mathematics, Princeton University, Princeton New Jersey 08544, USA (Dated: November 5, 2018) 2 1 0 2 n a J 4 ] h p - o i b . s c i s y h p [ 2 v 5 0 7 0 . 1 0 2 1 : v i X r a 1 Abstract Complex tumor-host interactions can significantly affect the growth dynamics and morphologies of progressing neoplasms. The growth of a confined solid tumor induces mechanical pressure and deformation of the surrounding microenvironment, which in turn influences tumor growth. In this paper, we generalize a recently developed cellular automaton model for invasive tumor growth in heterogeneous microenvironments [Y. Jiao and S. Torquato, PLoS Comput. Biol. 7, e1002314 (2011)] by incorporating the effects of pressure. Specifically, we explicitly model the pressure exerted on the growing tumor due to the deformation of the microenvironment and its effect on the local tumor-host interface instability. Both noninvasive-proliferative growth and invasive growth with individual cells that detach themselves from the primary tumor and migrate into the surrounding microenvironment are investigated. We find that while noninvasive tumors growing in "soft" homogeneous microenvironments develop almost isotropic shapes, both high pressure and host heterogeneity can strongly enhance malignant behavior, leading to finger-like protrusions of the tumor surface. Moreover, we show that individual invasive cells of an invasive tumor degrade the local extracellular matrix at the tumor-host interface, which diminishes the fingering growth of the primary tumor. The implications of our results for cancer diagnosis, prognosis and therapy are discussed. PACS numbers: 87.19.lx, 46.32.+x, 87.17.Pq 2 I. INTRODUCTION Tumor malignancy arises from many complex interactions occurring between the tumor and its host microenvironment.1,2 There is increasing evidence that the host microenviron- ment can significantly affect neoplasm progression.3 -- 20 The growth of a confined solid tumor also produces mechanical pressure, leading to deformation of the surrounding microenviron- ment, which generally affects the growth dynamics of the tumor.21 Such pressure can result in clinical complications, especially in a confined region of space such as the brain22 and deform or even collapse the intra-tumoral blood and lymphatic vessels.23 It has been hypothesized that pressure may also influence tumor physiology, growth rate and morphology.21 There- fore, understanding effects of pressure on tumor growth is important for both fundamental cancer research and clinical practice.21 Mechanical interactions between a tumor and its microenvironment is a topic of great interest. In the work of Helmlinger et al,21 it was shown that pressure can significantly reduce tumor growth rate and even inhibit tumor growth in vivo. Bru and Casero24 studied the effect of external pressure on the growth of tumor cell colonies and showed that tumor morphology strongly depends on the pressure exerted by the surrounding medium. Moreover, it has been observed that stiffer microenvironments can promote malignant behavior.25 For example, tumors embedded in low-density soft agarose gels remain roughly spherical in shape; however, they could exhibit a finger-like morphology in a stiff gel with high density.26 A variety of analytical and computational models have been developed to incorporate the effect of pressure on tumor growth. In particular, McElwain and Pettet27 considered that tumor cells mechanically behave as incompressible "bags" of fluid enclosed by the plasma membrane. Chen et al.28 modeled the growth of tumor spheroids in agarose gels, considering the agarose gel to be a elastic material undergoing large deformations and the tumor tissue was approximated by a fluid-like material with additional drag and surface tension effects. Roose et al.29 employed a linear poroelasticity model to estimate the solid stress generated by the growth of the tumor spheroid. Although the overall growth of the tumor spheroid can be well described by these analytical approaches, they are not able to provide detailed in- formation on the tumor morphology. Gevertz et al.7,8 employed a cellular automaton model to investigate the effects the shape of an organ on growing tumors through mechanical in- teractions. Using coupled nonlinear partial differential equations, Macklin and Lowengrub19 3 modeled the response of the tissue surrounding the tumor to the proliferation-induced me- chanical pressure. Specifically, these authors found that tumors growing in mechanically unresponsive (i.e., rigid) microenvironments develop invasive fingering morphologies and tumors growing in mechanically responsive (i.e., soft) microenvironment develop compact morphologies. However, since a continuum method is used, it is not possible to keep track of individual invasive cells that detach themselves from the primary tumor and how such invasive cells affect the growth dynamics and morphologies of the primary tumor. Recently, we presented a single-cell based cellular automaton (CA) model for invasive tumor growth in heterogeneous microenvironments30 in response to the challenge of devel- oping an "Ising" model for cancer growth.31 In this CA model, individual invasive cells can detach themselves from the primary tumor, locally degrade the extracellular matrix (ECM) and invade into the surrounding host microenvironment. A rich spectrum of emergent prop- erties and coupled growth dynamics of the primary tumor and invasive cells were predicted. However, the effects of pressure exerted by the outer boundary of the growth permitting region (e.g., cranium) on the tumor and the deformation of the ECM were only implicitly considered. In this paper, we generalize the aforementioned CA model to explicitly take into account the deformation of the ECM surrounding an invasive or noninvasive tumor, which in turn imposes pressure on the neoplasm. Moreover, we also explicitly consider the local geometry of the tumor-host interface (i.e., the tumor surface), which can either enhance or reduce local growth (i.e., interface instability) depending on the local curvature of the interface. Both noninvasive-proliferative growth and invasive growth with individual cells that detach themselves from the primary tumor and migrate into the surrounding microenvironment are investigated. We show here that by varying the ECM rigidity (density), one can obtain a continuous spectrum of tumor morphologies ranging from smooth isotropic shapes to finger- ing patterns, which have been observed both in vitro and in vivo19. The specific surface32 is employed to quantify the degree of "fingering" for noninvasive proliferative growth. We find that both the high pressure built up due to tumor growth and the microenvironment heterogeneity can significantly promote malignancy of the noninvasive proliferative tumor. Moreover, we show that individual invasive cells that leave an invasive primary tumor de- grade the local ECM at the tumor-host interface, which diminishes the fingering growth of the primary tumor. Our results concerning the diversity of tumor morphologies enable one 4 to infer what are the possible mechanisms behind the resulting shapes. Such information is expected to be of great value for cancer diagnosis, prognosis and therapy. II. COMPUTATIONAL METHODS Following Refs. [7,8,30,33 -- 35], we use the Voronoi tessellation associated with random- sequential-addition (RSA) sphere packings32 to model the underlying cellular structure. In particular, nonoverlapping d-dimensional spheres (d = 2 and 3) are randomly and sequen- tially placed in a prescribed region in d-dimensional Euclidean space until there is no void space left for additional spheres. Then, space is divided into polyhedra, each associated with a sphere center, such that any points within a polyhedron is closer to its associated sphere center than to any other sphere centers. The resulting Voronoi polyhedra are referred to as automaton cells, which can represent either real biological cells or regions of tumor stroma. Here we explicitly takes into account the interactions between a single cell and the sur- rounding microenvironment. Thus, each automaton cell represents either a single tumor cell (approximately 15 − 20 µm in size) or a region of tumor stroma of similar size. In the current model, we mainly focus on the effects of the ECM macromolecule density, ECM degradation by the malignant cells, and the pressure due to the ECM deformation on tumor growth. Henceforth, we will refer to the host microenvironment (or tumor stroma) as the "ECM" for simplicity. Each ECM associated automaton cell is assigned a particular density ρECM, representing the density of the ECM molecules within the automaton cell. A tumor cell can occupy an ECM associated automaton cell only if the density of this automaton cell ρECM = 0, which means that either the ECM is degraded or it is deformed (pushed away) by the proliferating tumor cells. A. Modeling the Pressure Exerted on the Growing Tumor The extracellular matrix is a complex mixture of macromolecules and interstitial fluids that provides mechanical supports for the tissue and plays an important role for cell adhesion and motility.36 In general, the ECM can be highly heterogeneous, with large spatial variations of the ECM macromolecule densities ρECM. Our simulated tumors are only allowed to grow in a compact growth-permitting region in order to mimic the physical confinement of the 5 host microenvironment, such as the boundary of an organ or cranium in the case of the brain. Therefore, a growing tumor deforms the ECM, which in turn imposes a pressure on the tumor.21,37,38 In Ref. 30, we considered the effects of the local ECM density on the proliferating cells. In this work, the ECM with larger density was considered to be more rigid and more difficult to degrade/deform. Therefore, the probability of division pdiv, which is related to the cell doubling time τ0 by τ0 = ln 2/ ln(1 + pdiv), was taken to be a monotonically decreasing function of ρECM, e.g., pdiv ∼ (1 − ρECM). (1) The effect of pressure was only considered implicitly, e.g., pdiv ∼ (1 − r/Lmax), where r is the distance of the dividing cell from the tumor centroid, Lmax is the distance between the closest growth-permitting boundary cell in the direction of tumor growth and the tumor centroid. Here we explicitly consider the pressure exerted on the growing tumor by the ECM due to deformation. In particular, we consider that the ECM is a linear elastic medium with bulk modulus κECM. The pressure P due to volume deformation ∆V is given by P = κECM ∆V V , (2) where V is the initial volume of the ECM. The ECM density ρECM = M/V , where M is the ECM mass. Therefore, the ECM density increase due to small volume shrinkage, i.e., ρECM = M/(V − ∆V ) ≈ (M/V )(1 + ∆V ) = ρ0 ECM(1 + ∆V ), which gives ∆V V = (ρECM − ρ0 ECM) ρ0 ECM . Substituting Eq. (4) into Eq. (2), we have P = κECM ∆V V = κECM (ρECM − ρ0 ECM) ρ0 ECM . (3) (4) (5) In other words, the pressure exerted by the ECM due to deformation is proportional to its density, i.e., P ∼ (ρECM − ρ0 ECM). Without loss of generality, we consider that cell divi- sion probability is simply a monotonically decreasing function of pressure, and thus, also a monotonically decreasing function of the ECM density, i.e., pdiv = (1 − ωP ) ∼ (cid:20)1 − ω κECM ρ0 ECM (ρECM − ρ0 ECM)(cid:21) = [1 − ω∗ ρ0 ECM (ρECM − ρ0 ECM)], (6) 6 where is a constant of proportionality. ω∗ = ωκECM (7) Suppose that the proliferative cells possess the ECM degradation ability χ0, which is the fraction of the ECM macromolecules degraded by malignant cells per day per unit volume. After each day, the total mass of the ECM that has been degraded is ∆M = χ0 n Xi ρECM(i)v(i), (8) where n is the total number of the ECM associated automaton cells taken by new tumor cells, ρECM(i) and v(i) are respectively the macromolecule density and volume associated with the ith automaton cell. The average ECM density is then given by ρECM = i ρECM(i)v(i) . M − χ0Pn V −Pn i v(i) We define the ratio of ρECM/ρ0 ECM to be ξ, i.e., ξ = ρECM ρ0 ECM = M − χ0Pn V −Pn i ρECM(i)v(i) i v(i) (9) (10) V M . The macromolecule densities of the remaining ECM automaton cells are then updated as ρECM(j) = ξρ0 ECM(j), (11) i.e., the increase of the ECM density after deformation is proportional to its original density. Substitute Eq. (11) into Eq. (6), we have pdiv ∼ [1 − ω∗(ξ − 1)] . (12) We note that in the above analysis, we have neglected the deformation of the tumor cells, which possess a much larger bulk modulus κcell than that of the ECM, i.e., κcell/κECM ∼ 100 (see Ref. 29). B. Modeling Local Tumor-Host Interface Instablity Real tumors never possess a perfect spherical shape. Tumors growing even in a homoge- neous soft ECM will develop a "bumpy" tumor surface, which can be very well captured by 7 our underlying cellular structure model (i.e., the Voronoi tessellation).30 When growing in a rigid microenvironment, a locally smooth tumor surface which results in a huge pressure gradient at the surface is highly undesirable. On the other hand, locally small protrusions on the tumor surface can gain some growth advantage by further invading into the surrounding ECM to release local pressure.26 To model the aforementioned effects, we consider the local geometry of the protrusion tip. In particular, the width of the tip is taken to be the length w of the automaton cell at the tip. The length of the tip is given by ℓ = xc − x, (13) where xc is the position of the center of the automaton cell at the tip and x = Pm i xi is the average center position of tumor cells neighboring the cell at the tip. The growth advantage of the cell at the tip is then proportional to ℓ/w, i.e., pdiv ∼ (1 + ℓ/w). (14) We note that ℓ is effective the radius of curvature, which can be either positive or negative. A negative value of ℓ reduces pdiv. For positive ℓ, the ratio ℓ/w is defined as the stress concentration factor associated with a crack tip in solid mechanics. Other biophysical mechanisms associated with noninvasive and invasive malignant cells are the same as those described in Ref. 30. For example, the non-invasive cells remain in the primary tumor and can be proliferative, quiescent or necrotic, depending on their nutritional supply, which is determined by tumor-size dependent characteristic diffusion distances. Proliferative cells can produce "mutant" daughter cells that possess strong ECM degradation ability χ1 and can leave the primary tumor and invade into the surrounding microenvironment by locally degrading the ECM macromolecules. Readers are referred to Ref. 30 for details of such mechanisms. C. Cellular Automaton Rules We now specify the CA rules for our generalized model, which closely follow those given in Ref. 30, except for the additional rules explicitly incorporating the pressure imposed by the ECM and the local host-tumor interface instability described here. After generating the 8 automaton cells using Voronoi tessellation, an ECM density ρECM ∈ (0, 1) is assigned to each automaton cell within the growth-permitting region, which represents the heterogeneous host microenvironment. Then a tumor is introduced by designating any one or more of the automaton cells as proliferative cancer cells. Time is then discretized into units that represent one real day. At each time step: • Each automaton cell is checked for type: invasive, proliferative, quiescent, necrotic or ECM associated. Invasive cells degrade and migrate into the ECM surrounding the tumor. Proliferative cells are actively dividing cancer cells, quiescent cancer cells are those that are alive, but do not have enough oxygen and nutrients to support cellular division and necrotic cells are dead cancer cells. • All tumorous necrotic cells are inert (i.e., they do not change type). • Quiescent cells more than a certain distance δn from the tumor's edge are turned necrotic. The tumor's edge, which is assumed to be the source of oxygen and nutrients, consists of all ECM associated automaton cells that border the neoplasm. The critical distance δn for quiescent cells to turn necrotic is computed as follows: δn = aL(d−1)/d t , (15) where a is a prescribed parameter (see Table I), d is the Euclidean spatial dimension and Lt is the distance between the geometric centroid xc of the tumor and the tumor edge cell that is closest to the quiescent cell under consideration. The position of the tumor centroid xc is given by xc = x1 + x2 + · · · + xN N , (16) where N is the total number of noninvasive cells contained in the tumor, which is updated when a new noninvasive daughter cell is added to the tumor. • Each proliferative cell will attempt to divide with probability pdiv into the surrounding ECM (i.e., the automaton cells associated with the ECM) by degrading and pushing away the ECM in that automaton cell. As discussed in the previous section, we consider that pdiv for a specific proliferative cell depends on the local ECM density 9 [Eq. (1)], the pressure imposed by the ECM [Eq. (6)] and the local geometry of the tumor-host interface [Eq. (14)], i.e., pdiv =   p0[1 − ρECM − ω∗ξ if any ECM associated automaton cell within +ω∗ + ξ ℓ w ] the predefined growth distance is in the growth- permitting microenvironment if no ECM associated automaton cell within (17) 0 the predefined growth distance is in the growth- permitting microenvironment, or the value of the above expression is negative where p0 is the base probability of division (see Table I), r is the distance of the dividing cell from the tumor centroid, ρECM is the ECM density of the automaton cell to be taken by the new tumor cell, ω∗ and ξ are respectively given by Eq. (7) and Eq. (10). When a ECM associated automaton cell is taken by a tumor cell, its density is set to be zero. The predefined growth distance (δp) is described in the following bullet point. • If a proliferative cell divides, it can produce a mutant daughter cell possessing an invasive phenotype with a prescribed probability γ (i.e., the mutation rate). The invasive daughter cell gains ECM degradation ability χ1 and motility µ, which enables it to leave the primary tumor and invade into the surrounding ECM. The rules for updating invasive cells are given in the following bullet point. If the daughter cell is noninvasive, it is designated as a new proliferative cell. • A proliferative cell turns quiescent if there is no space available for the placement of a daughter cell within a distance δp from the proliferative cell, which is given by δp = bL(d−1)/d t , where b is a nutritional parameter (see Table I), d is the spatial dimension and Lt is the distance between the geometric tumor centroid xc and the tumor edge cell that is closest to the proliferative cell under consideration. • An invasive cell degrades the surrounding ECM (i.e., those in the neighboring automa- ton cells of the invasive cell) and can move from one automaton cell to another if the 10 associated ECM is completely degraded locally. For an invasive cell with motility µ and ECM degradation ability χ1, it will make m attempts to degrade the ECM in the neighboring automaton cells and jump to these automaton cells, where m is an arbitrary integer in [0, µ]. For each attempt, the surrounding ECM density ρECM is decreased by δρ, where δρ is an arbitrary number in [0, χ1]. Using random numbers for the ECM degradation ability and cellular motility is to take into account tumor genome heterogeneity, which is manifested as heterogeneous phenotypes (such as dif- ferent m and δρ). When the ECM in multiple neighboring automaton cells of the invasive cell are completely degraded (i.e., ρECM = 0), the invasive cell moves in a direction that maximizes the nutrients and oxygen supply. Here we assume that the migrating invasive cells do not divide. The degraded ECM shows the invasive path of the tumor. • The density ρECM of the remaining ECM automaton cells is updated according to Eq. (11). The important parameters mentioned in the bullet points above are summarized in Table I. We note that although only spherical growth-permitting regions are considered here, this constraint can be easily relaxed. As a demonstration of the capability and versatility of the generalized CA model, we will employ it to investigate the growth dynamics and morphologies of both noninvasive and invasive tumors in two dimensions. However, the model is easily extended to three dimensions and the algorithmic details of the model are presented for any spatial dimension. III. RESULTS Homogeneous and random distributions of the ECM density30 are used to study the ef- fects of microenvironment heterogeneity on the growing tumor. The random distribution of the ECM density, which henceforth is referred to as the "random ECM" for simplicity, is generated by assigning a random ECM density value between 0 and 1 to each ECM as- sociated automaton cell. The boundary of the growth-permitting region is considered to be vascularized, i.e., a growing tumor can receive oxygen and nutrients from the growth- permitting region. In particular, we consider a constant radially symmetric nutrient/oxygen 11 TABLE I: Parameters and terms in the CA model. Summarized here are definitions of the pa- rameters for tumor growth and invasion, and all other (time-dependent) quantities used in the simulations. The number(s) listed in parentheses indicates the value or range of values assigned to the corresponding parameters in the simulations. The values of the parameters are chosen such that the CA model can reproduce reported growth dynamics of GBM from the medical literature3,7,33. Time dependent terms Lt Local tumor radius (varies with cell positions) Lmax Local maximum tumor extent (varies with cell positions) δp δn Characteristic proliferative rim thickness Characteristic living-cell rim thickness (determines necrotic fraction) pdiv Probability of division (varies with cell positions) ρECM ECM density (depends on ECM deformation and varies with positions, > 0) ξ ω∗ p0 a b ℓ Ratio of current ECM density over initial density ρECM/ρ0 ECM Parameter measuring pdiv reduced by pressure, = 2ρ0 ECM Growth parameters Base probability of division, linked to cell-doubling time (0.192) Base necrotic thickness, controlled by nutritional needs (0.58 mm1/2) Base proliferative thickness, controlled by nutritional needs (0.30 mm1/2) Length of local protrusion tip w Width of local protrusion tip χ0 γ χ1 µ ECM degradation ability of proliferative cells (0.0 − 0.25) Invasiveness parameters Mutation rate (determines the number of invasive cells, 0.05) ECM degradation ability of invasive cells (0.4 − 1.0) Cell motility (the number of "jumps" from one automaton cell to another, 0 − 4) 12 gradient in the growth-permitting region with the highest nutrient/oxygen concentration at the vascular boundary. We note that although generally the nutrient/oxygen concentration field in vivo is more complicated, previous numerical studies that considered the exact evolu- tion of nutrient/oxygen concentrations have shown a decay of the concentrations toward the tumor center.9,10 Since the directions of cell motions are determined by the nutrient/oxygen gradient only, our constant-gradient approximation is a very reasonable one. In the beginning, a proliferative tumor cell is introduced at the center of the growth- permitting region and tumor growth is initiated. The parameters employed are either given in Table I or specified for each case separately. The specific surface s for the noninvasive proliferative tumor, defined as the ratio of the total length of the perimeter of the primary tumor over its total area,30,32 is employed to quantify the degree of "fingering" of the growing tumor. For a perfectly circular shape with radius R, the associated s is given by 2/R, which is the minimal value among all shapes with the same area. The specific surface of a tumor in excess of that of a circle provides a measurement of the roughness of the tumor surface and thus, the degree of "fingering". Therefore, the specific surface s scaled by 2/RT associated with a circle is used for an arbitrary-shaped tumor with effective radius RT (i.e., the average distance from tumor edge to tumor center). In the visualizations of the tumor that follow, we will use the following convention: The ECM degraded by the tumor cells is blue. In the primary tumor, necrotic cells are black, quiescent cells are yellow and proliferative cells are red. The invasive tumor cells are green. A. Noninvasive Proliferative Growth We first investigate the effects of pressure on the growth of noninvasive proliferative tumors by setting the mutation rate to zero, i.e., γ = 0. Figure 1 shows the snapshots of noninvasive tumors growing in the homogeneous ECM with initial density ρ0 ECM = 0.25, 0.45 and 0.65. The associated specific surface as a function of time is also shown. The ECM degradation ability value χ0 = 0.2 for the proliferative cells is used. It can be clearly seen that as the ECM density (i.e., pressure level) varies, a variety of growth dynamics and tumor morphologies emerge. In particular, the tumor growing in the ECM with low density (pressure) (ρECM = 0.25) develops an almost isotropic shape with small specific surface s/(2/RT ) (upper panel of Fig. 1). As the ECM density (pressure) increases 13 (ρECM = 0.45), the tumor begins to develop bumpy surface, but a well-defined "spherical" core can still be identified (middle panel of Fig. 1). For the high ECM density (pressure) ρECM = 0.65, finger-like protrusions emerge at early growing stages and no "spherical" core of significant size is found (lower panel of Fig. 1). The associated specific surface s/(2/RT ) monotonically increases as the tumor grows. Since the elongated finger-like structures of proliferative cells can grow into the nutrient-rich microenvironment, necrotic regions within these "fingers" are very rare. We note that the overall size (i.e., distance between tumor center and the farthest cell on tumor edge) of the tumors with distinct morphologies decrease as the ECM density and the associated pressure increase. Nonetheless, the emergence of finger-like structures significantly release the local pressure built up at the smooth surface of the growing tumor and thus, the growth of such invasion fingers is favored. Figure 2 shows the snapshots of noninvasive tumors (e.g., γ = 0) growing in the random ECM with average initial density ρ0 ECM = 0.25, 0.45 and 0.65, as well as the associated specific surface s/(2/RT ) as a function of time. The ECM degradation ability value χ0 = 0.2 for the proliferative cells is used. Similar effects of the ECM density (i.e., pressure level) on the growth dynamics and tumor morphologies are observed, i.e., increasing the ECM density (pressure) leads to a continuous variation of tumor morphology ranging from smooth isotropic shapes to significantly fingered patterns. Since the heterogeneity of the ECM results in stronger local tumor surface roughness, the fingering effect is also stronger comparing with the corresponding homogeneous case with the same ECM density. In particular, the tumor growing in the ECM with ρ0 ECM = 0.65 develops sub-fingers on the primary fingers, which is mainly caused by the local ECM density fluctuations (lower panel of Fig. 2). Such fine morphological features can hardly be resolved by continuum simulation method. The stronger fingering of tumors growing in the random ECM also leads to larger overall extents of the tumors. B. Invasive Growth with Individual Cells Migrating into Surround ECM Figure 3 shows the snapshots of invasive tumors growing in a homogeneous ECM with initial density ρ0 ECM = 0.25, 0.45 and 0.65 on day 120. Specifically, individual invasive cells can detach themselves from the primary tumor and migrate into the surrounding ECM. The following values of invasiveness parameters are used: mutation rate γ = 0.05, ECM 14 degradation ability value for the proliferative cells χ0 = 0.25, ECM degradation ability value for the invasive cells χ1 = 0.75, motility of the invasive cells µ = 4, which corresponds to a high degree of malignancy. It is clear from Fig. 3 that the tumor growing in the low-density ECM (e.g., under low pressure) (ρECM = 0.25) develops relatively short invasive branches (Fig. 3a), while the tumor growing in the high-density ECM (e.g., under high pressure) (ρECM = 0.65) possesses very long invasive branches (Fig. 3c). Also, we note that a large number of invasive cells are generated at the finger tips of the primary tumor, which in turns promotes the growth of the fingers. Since the invasive cells degrade the ECM close to the tumor surface, the local pressure exerted on the tumor is reduced. Thus, the degree of "fingering" in the invasive tumors is smaller than in the noninvasive ones. The invasive cells also enhance the growth of the primary tumor. Moreover, it can be seen that high ECM density (i.e., high pressure exerted on the tumor) enhances both fingering of the primary tumor and malignant behavior of invasive cells (Fig. 3c). We note that such invasion-tumor couplings have been extensively explored in Ref. 30. Figure 4 shows the snapshots of invasive tumors growing in the random ECM with average initial density ρ0 ECM = 0.25, 0.45 and 0.65 on day 120. The same values of invasiveness parameters as the homogeneous case are used. Again, the growth dynamics and tumor morphologies are qualitatively the same as the corresponding homogeneous case. However, the heterogeneity of the ECM enhances local tumor instability and thus, promotes the malignant behavior of the invasive tumor. IV. CONCLUSIONS AND DISCUSSION We have generalized a recently developed cellular automaton model to study the effects of pressure on the dynamics and tumor morphologies for both noninvasive proliferative growth and invasive growth with individual malignant cells detaching themselves from the primary tumor. In particular, we have explicitly taken into account the deformation of the extracellular matrix surrounding the tumor, which in turn imposes pressure on the neoplasm. Moreover, we also considered the local tumor-host interface instability, which can give rise to the emergence of finger-like protrusions of the tumor surface. We showed that by varying the ECM rigidity (i.e., the pressure level in the ECM), one can obtain a 15 variety of tumor morphologies ranging from spherical shapes to fingering patterns, which are quantitatively characterized by the specific surface. We also found that both high pressure and microenvironment heterogeneity can amplify the malignancy of the neoplasm in both the noninvasive proliferative case and the invasive case. Moreover, we demonstrated that the growth dynamics of the primary tumor and invasive cells are strongly coupled. Figure 5 shows the morphology of ductal carcinoma in situ (DCIS), which resembles the morphology of noninvasive tumors growing in the ECM with intermediate density (Fig. 1(b) and Fig. 2(b)) predicted by our CA model. Strong fingering growth is rarely observed for DCIS, manly because the neoplasm is further constrained by a tight basal membrane composed of epithelial cells. Such complex microenvironment heterogeneities need to be incorporated in the CA model to accurately predict DCIS progression. On the other hand, significant fingering has been observed in vitro.19 These experimental observations clearly demonstrate the robustness and predictive capability of our CA model. We have shown that a high pressure in the ECM, which is due to large ECM density and deformation, can lead to significant fingering growth and enhance the malignant behavior of both the primary tumor and invasive cells. As a tumor grows in a confined microen- vironment, it is inevitable that a high pressure will be built up. Thus, a tumor with a low level of malignancy initially can eventually develop highly malignant invasive behavior. Specifically, when finger-like protrusions develop, cells close to the finger tip have less con- tacting neighbors and thus, less adhesion. This makes it very easy for the invasive cells to leave the primary tumor and migrate deeply into the surrounding microenvironment, which ultimately leads to cancer metastasis. Moreover, our results concerning the diversity of tumor morphologies enable one to infer the possible mechanisms behind the resulting shapes. For example, a noninvasive prolif- erative tumor possesses a morphology with significant finger-like protrusions on the tumor surface could be attributed to a host microenvironment in which the tumor grew that was very rigid and inhomogeneous. On the other hand, if such a tumor has a smooth and almost isotropic shape, it could mean that its host environment was very soft and homogeneous. For an invasive tumor with individual invasive cells that detach themselves from the primary tumor and migrate into the surrounding microenvironment, a rougher tumor surface could imply that the individual invasive cells possessed a strong ECM degradation ability, high motility and weak cell-cell adhesion. 16 Although our CA model is readily applied to model in vitro tumor growth, the hetero- geneous microenvironments considered in the current model are highly idealized and do not include heterogeneities such as blood vessels and lymphatics, which could play an important role in clinical cancers. Incorporating more realistic microenvironments as well as other pos- sible mechanisms (e.g., tumor and normal cell phenotypic plasticity and immune response) would lead to an improved model that could provide insights into in vivo tumor growth. Nonetheless, we expect that the conclusions drawn here still qualitatively apply to in vivo situations. Therefore, information on the tumor morphology, which can be obtained from histological images, is expected to lead to more accurate diagnosis and thus, more effec- tive tumor treatment strategies. For example, if a tumor with a rough surface is detected, drugs that can release the high concomitant pressure in the host environment by modify- ing the molecular compositions of the ECM macromolecules could be used to reduce the malignancy of the tumor, leading to a noninvasive smooth isotropic shape. This not only could improve the efficiency of chemotherapy but also could make it easier to remove the tumor by resection. Finally, we note that improving the deliverability of chemotherapy will require application of existing theories to predict the transport properties of the underlying heterogeneous media.39 Acknowledgments The research described was supported by the National Cancer Institute under Award NO. U54CA143803. The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Cancer Institute or the National Institutes of Health. ∗ Electronic address: [email protected] † Electronic address: [email protected] 1 D. S. Coffey DS, Nat. Med. 4, 882 (1998). 2 D. Hanahan and R. A. Weinberg, Cell 100, 57 (2000). 3 T. S. Deisboeck, M. E. Berens, A. R. Kansal, S. Torquato, A. Rachamimov, D. N. Louis and E. A. Chiocca, Cell Proliferat. 34, 115 (2001). 17 4 I. J. Fidler, Nat. Rev. Cancer. 3, 435 (2003). 5 A. R. A. Anderson and M. A. J. Chaplain, Bull. Math. Biol. 60, 857 (1998). 6 J. L. Gevertz and S. Torquato, J. Theor. Biol. 243, 517 (2006). 7 J. L. Gevertz, G. T. Gillies and S. Torquato, Phys. Biol. 5, 036010 (2008). 8 J. L. Gevertz and S. Torquato, Phys. Rev. E 80, 051910 (2009). 9 A. R. A. Anderson, Math. Med. Biol. 22, 163 (2005). 10 A. R. A. Anderson, A. M. Weaver, P. T. Cummings and V. Quaranta, Cell 127, 905 (2005). 11 R. A. Gatenby and E. T. Gawlinski, Cancer Res. 56, 5745 (1996). 12 R. A. Gatenby, E. T. Gawlinski, A. F. Gmitro, B. Kaylor and R. J. Gillies, Cancer Res. 66, 5216 (2006). 13 H. B. Frieboes, X. M. Zheng, C. H. Sun, B. Tromberg, R. Gatenby and V. Gristini, Cancer Res. 66, 1597 (2006). 14 N. Bellomo and L. Preziosi, Math. Comput. Model. 32, 413 (2000). 15 M. Scalerandi, B. C. Sansone and C. A. Condat, Phys. Rev. E 65, 011902 (2001. 16 M. Scalerandi and B. C. Sansone, Phys. Rev. Lett. 89, 218101 (2002). 17 Y. Kim and A. Friedman, Bull. Math. Biol. 72, 1029 (2010). 18 A. M. Stein, T. Demuth, D. Mobley and M. Berens, Biophys. J. 92, 356 (2007). 19 P. Macklin and J. Lowengrub, J. Thero. Biol. 245, 677 (2007). 20 Y. Jiao and S. Torquato, PLoS One 6, e27323 (2011). 21 G. Helmlinger, P. A. Netti, H. C. Lichtenbeld, R. J. Melder and R. K. Jain, Nature Biotech. 15, 778 (1997). 22 M. C. Hogan, A. Lee, L. A. Solberg and S. D. Thome, Am. J. Hematol. 70, 55 (2002) 23 T. P. Padera, A. Kadambi, E. di Tomaso, C. M. Carreira, E. B. Brown, Y. Boucher, N. C. Choi, D. Mathisen, J. Wain, E. J. Mark, L. L. Munn and R. K. Jain, Science 296. 1883 (2002). 24 A. Bru and D. Casero, J. Thero. Biol. 243, 171 (2006). 25 M. J. Paszek, N. Zahir, K. R. Johnson, J. N. Lakins, G. I. Rozenberg, A. Gefen, C. A. Reinhart- King, S. S. Margulies, M. Dembo, D. Boettiger, D. A. Hammer and V. M. Weaver, Cancer. Cell 8, 241 (2005). 26 C. Guiot, P. Nicola, P. P. Delsanto and T. S. Deisboeck, Phys. Biol. 4, P1 (2007). 27 D. L. McElwain and G. J. Pettet, Bull. Math. Biol. 55, 655 (1993). 28 C. Y. Chen, H. M. Byrne and J. R. King, J. Math. Biol. 43, 191 (2001). 18 29 T. Roose, P.A. Netti, L.L. Munn, Y. Boucher and R.K. Jain, Microvascular Res. 66, 204 (2003). 30 Y. Jiao and S. Torquato, PLoS Comput. Biol. 7, e1002314 (2011). 31 S. Torquato, Phys. Biol. 8, 015017 (2011). 32 S. Torquato, Random Heterogeneous Materials: Microstructure and Macroscopic Properties. (Springer-Verlag, New York, 2002). 33 A. R. Kansal, S. Torquato, G. R. Harsh, E. A. Chiocca and T. S. Deisboeck, J. Theor. Biol. 203, 367 (2000). 34 A. R. Kansal, S. Torquato, E. A. Chiocca and T. S. Deisboeck, J. Theor. Biol. 207, 431 (2000). 35 J. E. Schmitz, A. R. Kansal and S. Torquato, J. Theor. Med. 4, 223 (2002). 36 L. A. Liotta, C. N. Rao and S. H. Barsky, Lab. Invest. 49, 636 (1983). 37 M. Sarntinoranont, F. Rooney and M. Ferrari, Annals of Biomedical Engineering 31, 327 (2003). 38 V. D. Gordon, M. T. Valentine, M. L. Gardel, D. Andor-Ardo, S. Dennison, A. A. Bogdanov, D. A. Weitz and T. S. Deisboeck, Experimental Cell Res. 289, 58 (2003). 39 S. Torquato, J. Chem. Phys. 84, 6345 (1986); J. Rubinstein and S. Torquato, J. Fluid Mech. 206, 25 (1989); S. Torquato and A. K. Sen, J. Appl. Phys. 67, 1145 (1990); S. Torquato, Phys. Rev. Lett. 64, 2644 (1990); S. Torquato and M. Avellaneda, J. Chem. Phys. 95, 6477 (1991). 19 FIG. 1: Upper panel: Snapshots of a noninvasive tumor growing in homogeneous ECM with density ρECM = 0.25 on day 40 (a1), day 80 (a2) and day 120 (a3) after initialization. The associated specific surface s/(2/RT ) (a4) remains small in value as the tumor grows, indicating a compact tumor morphology. Middle panel: Snapshots of a noninvasive tumor growing in homogeneous ECM with density ρECM = 0.45 on day 40 (b1), day 80 (b2) and day 120 (b3) after initialization. The associated specific surface s/(2/RT ) (b4) shows a rapid growth after day 100, indicating fingering growth of the tumor. Lower panel: Snapshots of a noninvasive tumor growing in homogeneous ECM with density ρECM = 0.65 on day 40 (c1), day 80 (c2) and day 120 (c3) after initialization. The associated specific surface s/(2/RT ) (c4) increase monotonically as the tumor grows, indicating significant fingering of the tumor. FIG. 2: Upper panel: Snapshots of a noninvasive tumor growing in random ECM with average density ρECM = 0.25 on day 40 (a1), day 80 (a2) and day 120 (a3) after initialization. The associated specific surface s/(2/RT ) (a4) remains small in value as the tumor grows, indicating a compact tumor morphology. Middle panel: Snapshots of a noninvasive tumor growing in random ECM with average density ρECM = 0.45 on day 40 (b1), day 80 (b2) and day 120 (b3) after initialization. The associated specific surface s/(2/RT ) (b4) shows a rapid growth after day 100, indicating fingering growth of the tumor. Lower panel: Snapshots of a noninvasive tumor growing in random ECM with average density ρECM = 0.65 on day 40 (c1), day 80 (c2) and day 120 (c3) after initialization. The associated specific surface s/(2/RT ) (c4) increase monotonically as the tumor grows, indicating significant fingering of the tumor. Note that sub-fingers are developed on the primary fingers. FIG. 3: Snapshots of invasive tumors growing in a homogeneous ECM with different densities on day 100. Individual invasive cells detach themselves from the primary tumor, locally degrade the ECM and migrate into the surrounding microenvironment. (a) ρECM = 0.25 (b) ρECM = 0.45 (c) ρECM = 0.65. Note that the tumor growing in the low-density ECM (pressure) (ρECM = 0.25) develops relative short invasive branches [see panel (a)], while the tumor growing in the high-density ECM (pressure) (ρECM = 0.65) possesses very long invasive branches [see panel (c)]. 20 FIG. 4: Snapshots of invasive tumors growing in a random ECM with different average densities on day 100. Individual invasive cells detach themselves from the primary tumor, locally degrade the ECM and migrate into the surrounding microenvironment. (a) ρECM = 0.25 (b) ρECM = 0.45 (c) ρECM = 0.65. Note that the tumor growing in the low-density ECM (pressure) (ρECM = 0.25) develops relative short invasive branches [see panel (a)], while the tumor growing in the high- density ECM (pressure) (ρECM = 0.65) possesses very long invasive branches [see panel (c)]. Also observe that invasive cells clump at the finger tips of the primary tumor, which in turns promotes the growth of the fingers [see panel (c)]. FIG. 5: An image showing the morphology of ductal carcinoma in situ with bumpy surface. Image courtesy of R. Gatenby. 21 (a1) (a2) (a3) 4 3 2 s 1 0 0 5 4 3 s 2 1 0 0 40 t (days) 80 120 (a4) 40 t (days) 80 120 (b1) (b2) (b3) (b4) 10 s 5 0 0 40 t (days) 80 120 (c1) (c2) (c3) (c4) FIG. 1: Jiao, Torquato 22 (a1) (a2) (a3) (b1) (b2) (b3) 4 3 2 s 1 0 0 7 6 5 4 s 3 2 1 0 0 20 15 s 10 5 0 0 40 t (days) 80 120 (a4) 40 t (days) 80 120 (b4) 40 t (days) 80 120 (c1) (c2) (c3) (c4) FIG. 2: Jiao, Torquato 23 (a) (b) (c) FIG. 3: Jiao, Torquato (a) (b) (c) FIG. 4: Jiao, Torquato FIG. 5: Jiao, Torquato 24
1202.3401
1
1202
2012-02-15T18:32:47
Role of disclinations in determining the morphology of deformable fluid interfaces
[ "physics.bio-ph", "cond-mat.soft", "physics.comp-ph" ]
We study the equilibrium shapes of vesicles, with an in-plane nematic order, using a Monte-Carlo scheme and show that highly curved shapes, like tubes and discs, with a striking similarity to the structures engendered by certain curvature sensing peripheral membrane proteins, can be spontaneously generated by anisotropic directional curvature with nematic disclinations playing and important role. We show that the coupling between nematic order and local curvature could lead to like defects moving towards each other and unlike defects moving away, in turn leading to tube formation. Thermally induced defect pair production lead to branched tubular structures. It is also shown that helical arrangement of the membrane tubes, with nematic field spiraling around it, is a dominant soft mode of the system.
physics.bio-ph
physics
Role of disclinations in determining the morphology of deformable fluid DOI: 10.1039/c2sm07384f, Soft Matter Communications interfaces N. Ramakrishnan∗ and P. B. Sunil Kumar† Department of Physics, Indian Institute of Technology Madras, Chennai 600036, India John H. Ipsen‡ MEMPHYS- Center for Biomembrane Physics, Department of Physics and Chemistry, University of Southern Denmark, Campusvej 55, DK-5230 Odense M, Denmark (Dated: November 11, 2018) Abstract We study the equilibrium shapes of vesicles, with an in-plane nematic order, using a Monte-Carlo scheme and show that highly curved shapes, like tubes and discs, with a striking similarity to the structures engendered by certain curvature sensing peripheral membrane proteins, can be spontaneously generated by anisotropic directional curvature with nematic disclinations playing and important role. We show that the coupling between nematic order and local curvature could lead to like defects moving towards each other and unlike defects moving away, in turn leading to tube formation. Thermally induced defect pair production lead to branched tubular structures. It is also shown that helical arrangement of the membrane tubes, with nematic field spiraling around it, is a dominant soft mode of the system. PACS numbers: PACS-87.16.D-, Membranes, bilayers and vesicles. PACS-05.40.-a Fluctuation phenomena, random pro- cesses, noise and Brownian motion. PACS-05.70.Np Interfaces and surface thermodynamics 2 1 0 2 b e F 5 1 ] h p - o i b . s c i s y h p [ 1 v 1 0 4 3 . 2 0 2 1 : v i X r a ∗Electronic address: [email protected] †Electronic address: [email protected] ‡Electronic address: [email protected] 1 Fluid interfaces with an inplane orientational order can sustain amazingly complex morphologies. Apart from the interesting physics that it can offer, in terms of disclination dynamics on deformable surfaces, study of such interfaces are also contribute to understanding morphologies of cellular organelles. The organelles of a biological cell have membranes with highly curved edges and tubes, as seen in the Endoplasmic reticulum, the Golgi and the inner membrane of mitochondria. Tubulation has also been observed, in vitro, in self assembled systems of pure lipids [1]. It has been shown that, macro- molecules, which constitutes and decorates the membrane surface, strongly influence the morphology of membranes. For instance proteins from the dynamin superfamily are known to pull out membrane tubes while oligomerizing themselves into a helical coat along the tube[2]. The BAR domain containing proteins in general can induce a wide spectrum of membrane shapes ranging from protrusions to invaginations depending on the geometry and interaction strength of the BAR domain [3 -- 5]. The standard Helfrich model[6] for membranes, based on mean curvature energy, cannot explain the stability of such highly curved structures. Existence of anisotropic bending energy will be the minimal requirement to explain the stability of tubular shapes, which could arise from an in-plane orientational field on the membrane [7, 8]. Planar orientational order could be intrinsic to the membrane [9], due to the structural properties of its constituents, or could arise as a result of membrane interactions with (cid:48) external agents [3 -- 5]. Lipid molecules tilted with respect to the layer normal, as seen in the L β [10] (cid:48) and P β [11] phases, is a well known case of intrinsic in-plane orientational order. Anisotropic peripheral proteins (curvactants), inducing membrane deformations [12] either by scaffolding or by insertion of their amphipathic helices into the membrane [3], are examples for extrinsic sources of orientational order. In general the order can be represented by a p-atic surface field, invariant under rotations by 2π/p. Equilibrium shapes of a surface with p-atic in plane order, have been investigated earlier [13, 14]. Mean field phenomenological models of orientation order on membrane surface, in the context of lipid tilt and chirality, have predicted the stability of tubular membranes and helical ribbons[15 -- 18]. More recently, a mesoscopic particle model was used to show that presence of BAR domains on membrane surfaces can lead to tubulation and vesiculation [19, 20]. The coupling of curvature to orientational order leads to interesting defect dynamics and thermally excited shapes. Thus in addition to its biological relevance, membranes with in plane order also provides a unique medium to study disclination dynamics on deformable surfaces. So far such studies have been carried out only on surfaces with prescribed curvature [21]. Apart from the interesting physics that it can generate these investigations may also provide new routes to new functional materials[22]. In this paper we use the Monte Carlo model developed by us [23, 24] to explore the spectrum of vesicle shapes engendered by an in-plane nematic field m. The technique developed by us enables one to go beyond the mean field regime and also to explore the full nonlinear regime. We scale all energies by kBT . A moderate mean curvature stiffness κ = 10 is chosen to enable study of thermally induced shape fluctuations. Higher values of κ are chosen to study equilibrium shapes in the absence of thermal fluctuations. We choose the directional curvature moduli, along and perpendicular to the orientational field, as κ(cid:107) = 5 and κ⊥ = 0 respectively. The results with κ⊥ (cid:54)= 0 is qualitatively same and is not discussed separately. We vary the spontaneous curvature along the in-plane field C(cid:107)0 , to explore different shapes. The in-plane nematic field, distributed uniformly over the entire surface, is constrained to be in the ordered phase by setting the Lebwohl Lasher coupling constant LL = 3kBT [25]. For this values of the coupling constant, with κ(cid:107) = 0, the nematic field orients itself on a tetrahedron with four +1/2 disclinations, as predicted in [13]. We will first investigate the effect of the directional spontaneous curvature on the equilibrium membrane conformations when the κ value is high so that thermal undulations can be neglected. At high κ thermal undulation does not change the shapes significantly and the shapes obtained from the equilibrium simulations will be the same as that from a mean field calculation. It is known that dynamic triangulation Monte Carlo (DTMC) models, based on Helfrich Hamiltonian [6], do reproduce the mean field phase diagram of vesicles [26]. Here we determine the equilibrium shapes of decorated vesicles with an in-plane orientational order, also taking thermal fluctuations into account . As can be seen from Fig.1, we see tubular and disc like shapes emerging, even when there are no constraints on the area to volume ratio of the vesicle. Directional spontaneous curvature and interaction between the disclinations are two important factors that determine the shape of these decorated vesicles. Since the membrane is self avoiding the constraint imposed by the fixed topology also plays a role. Four curvatures can be identified, the curvatures along the principal directions c1, c2 with c1 > c2 and the directional curvatures, c(cid:107), c⊥, along and perpendicular to m. Fig.2 shows the distribution P (ci), of 2 FIG. 1: Membrane conformations for different values of C(cid:107)0 with κ = 10, κ(cid:107) = 5, κ⊥ = 0 and LL = 3. ( See supplementary materials for high resolution figures of nematic field arrangement.) FIG. 2: (Color online) (a,b) Distribution of the maximum and minimum principal curvatures, c1 and c2 respectively, for different values of C(cid:107)0 . ci=1,2, on a vesicle, for different C(cid:107)0 values when κ⊥ = 0. It is clear from Fig.2(a & b) that for C(cid:107)0 > 0 the distributions have a single peak, with P (c1), peaked close to, but less than, the value of C(cid:107)0 , while P (c2) is peaked at zero. The resultant value of the principal curvature arising from the competition between the resistance to bending imposed by κ and the directional spontaneous curvature set by C(cid:107)0 , is shown in Fig.3. When κ decreases the peak position of P (c1) moves toward that of C(cid:107)0 and then to a value higher than C(cid:107)0 for κ ∼ 0. Note that for κ = 0 there are no competing elastic forces and the desired directional curvature can be achieved by setting m at an angle ϕ (cid:54)= 0 with respect to the direction of c1. For a tube, a non zero value of ϕ implies a configuration with nematic spiraling around the tube. The angle ϕ is shown as inset in Fig. 3. As expected this angle is a decreasing function of the bending modulus κ and 3 (a) Prolate(b) Tubular(c) Disc(d) Caveola like(e) Tubular(f) Spirals(g) Narrow neck(h) Broad neck-0.6-0.4-0.20.00.20.40.50.600.0050.01-1-0.500.51-1-0.500.5100.0050.010.20.50.7-0.2-0.5-0.7(a)(b)c1&c2P(c1)P(c2) saturates to zero for large bending rigidity. Negative spontaneous directional curvature (C(cid:107)0 < 0) induces tubes to form into the vesicle. However, restrictions imposed by the membrane self avoidance and topology prevent a complete inversion of the membrane through inward tubulation. This leads to a bimodal distribution of c2, with a peak at c2 = 0 in addition to the expected peak at c2 ≈ C(cid:107)0 (see Fig.2 a). This bimodal peak arises due to presence of inward tubes (with most of the vertices with c2 < 0, c1 = 0 ) and an outward curved surface (with most vertices having c1 > 0, c2 = 0). FIG. 3: (Color online) Comparison of the c1 (filled symbols) and directional curvature c(cid:107) (open symbols) distribution, at C(cid:107)0 = 0.4, κ⊥ = 0, for different values of κ. Inset shows the change in the angle ϕ as a function of κ. Dotted vertical line marks c1 = C(cid:107)0 = 0.4. Another important factor that determines the shape of a vesicle is the positioning of disclinations. We know that, for a nematic field on a closed surface with spherical topology, the total disclination strength should be 2. There is no such topological restriction on the number of defects themselves and thermal fluctuation could excite oppositely charged defect pairs. FIG. 4: (Color online) Distribution of the nematic defect separation (ξ) for κ = 10, LL = 3 and κ⊥ = 0. In the absence of any stiffness-orientation coupling, the energy of a pair of defects, on a surface of spherical topology, decreases logarithmically with separation [27] and is repulsive when their charges are of the same sign [13]. We therefore expect, at low temperatures, to see four +1/2 disclinations at equal distance from each other. In Fig.4 this is evidenced in the single peak in the measured distribution of geodesic distances (P (ξ)) between the defects. Directional bending stiffness ( κ(cid:107) (cid:54)= 0 ) alters this distribution considerably, since curvature influences the interaction between defects [28, 29]. We see that two of the +1/2 defects move towards each other to form pairs as the vesicle deforms into a tube, resulting in a peak in the ξ distribution at short separation as shown in Fig.4. This peak moves towards the left as C(cid:107)0 increases. The pairs, which now has a total strength of +1, moves away from each other, resulting 4 00.511.500.0030.0060.009051015200.40.60.81κ=0κ=5κ=10κhcosϕick,c1P(ck,c1)filled(c1)open(ck)Ck0=0.3Ck0=0.4Ck0=0.5025507510012500.010.020.030.040.05separation between defects pairs.Defects at equal distancereplacemenGeodesicdistance,ξP(ξ)κk=0;Ck0=0κk=5;Ck0=0.2κk=5;Ck0=0.3κk=5;Ck0=0.4κk=5;Ck0=0 FIG. 5: (Color online) Line discontinuities in the nematic orientation for C(cid:107)0 < 0. (a) Paired tubes have rims that end on −1/2 defects while (b) closed rims are surrounded by four −1/2 defects (b). in a second peak in P (ξ). This peak position moves to the right as C(cid:107)0 is increased. We thus observe nontrivial changes in the interaction between disclinations due to the geometry of the embedding surface. For C(cid:107)0 < 0, there is a proliferation of defects as the inward tubes are formed. The inner side of the tubes are decorated with m along the circumference with the tip hosting a pair of +1/2 defects. There is a sudden change in the orientation of m just out side the rim of the tube(see Fig.5). The rim itself is thus a line of discontinuity. There are two types of rims, closed ones with four −1/2 defects outside and paired open rims ending in two −1/2 defects. A variety of thermally excited shapes are also seen. The lowest energy thermal modes among the tubes are the spirals, which are always present unless one is at a very low temperature (see figure 1f). If we raise the temperature or decrease the value of LL, the tubes develop branches. More branches are produced as the temperature is increased, with every branch having a pair of −1/2 defects at the intersection and a pair of +1/2 defects at the tip [24]. Fig.6 shows the sequence of snap shots that depict the transformation of a sphere with an initial random orientation of the nematic to a tube, when quenched from LL = 0 to LL = 3, while keeping κ(cid:107) = 5, κ⊥ = 0.0, κ = 10 and C(cid:107)0 = 0.4 [30]. The nematic instantly orients on the sphere resulting in four +1/2 defects at equal distance from each other. The membrane then begins to deform, through formation of protrusions at the defect cores (Fig.6(a)), since positive defects favor positively curved regions[28] and to satisfy the curvature set by C(cid:107)0 and κ. These membrane protrusions are accompanied by production of oppositely charged half defect pairs(Fig.6(b)). The negative defects move to the negatively curved intersections and the positive defects move to the positively curved tip of the protrusions. When the tip of a protrusion has two +1/2 defects, and its neck two −1/2, that branch is stabilized and grows by moving the positive and negative defect pairs away from each other(Fig.6(c)). Isolated +1/2 and −1/2 defects annihilate(Fig.6(d)). In general two types of branched structures are observed, as we increase the directional spontaneous curvature, (a) Broad necked, see Fig.1(g) and (b) Narrow necked, see Fig.1(h) . (cid:18) To analyze the origin of helicity, we neglect the end cap effects of the tube and parame- is then [24] Htot = 0 r)(1 − 2x − C⊥ π 0 r) r Here r is the radius of the tube and x = cos2 ϕ. The energy depends only on the curvature of its (cid:18) κ+κ(cid:107)(1−x)2+κ⊥x2 1 −(cid:112)1 − (λr)2 (cid:16) κ(cid:107) 2 (1 − C The total elastic energy, per unit length, (cid:107) 0 r) + κ⊥ 0 r)(2x − 1 − HC (cid:107) + √ terize it as a canal surface. (cid:16) LL 2 (1 − C⊥ (cid:17) + (cid:19) 1−(λr)2 (cid:17)(cid:19) . 5 FIG. 6: (Color online) The formation of tubes and branches are driven by proliferation of a +1/2 and -1/2 defect pairs . See text for details. spine curve λ, and not on the torsion, hence for λ (cid:54)= 0 there is a degeneracy of spiral configurations corresponding to different torsions. One can easily see that the energy is at its minimum when λ = 0 [24]. For a tube of length L, parametrizing the spiral deformations by the extension of the tube along the helical axis Lz = L√1 − λr, we can estimate the entropy to be of the order kBT ln( L a ) and for small λ values the entropy term will dominate the free energy resulting in helical configurations. Before we conclude, let us look at how does the model parameters like the spontaneous curvature and is about 4 tether lengths. Comparing this with the radius of the lipid tubes obtained in experiments [31] the density of nematic inclusions compare with experiments. For C(cid:107)0 ∼ 1.0, the circumference of the tube we get the length of the tether to be ≈ 25nm. The value of C(cid:107)0 = 1.0 in real units is thus ≈ (25 nm)−1, which is not far from the suggested value of dynamin intrinsic curvature. Experiments see the dynamin rings surrounding the tubes to be made of about 20 units with a pitch of about 15 nm. This will translate to about 5 dynamin molecules per vertex in the simulations. In conclusion, we have shown that in-plane nematic order couple to curvature, on a deformable surface, can lead to non-trivial shapes of vesicles. The deformability of the surface lead to generation of point defects and line singularities,which in turn leads to production of tubes and branches. [1] M. Markowitz and A. Singh, Langmuir 7, 16 (1991). [2] G. J. K. Praefcke and H. T. McMahon, Nat. Rev. Mol. Cell Biol. 5, 133 (2004). [3] J. Zimmerberg and M. M. Kozlov, Nat. Rev. Mol. Cell Biol. 7, 9 (2006). [4] G. K. Voeltz and W. A. Prinz, Nat. Rev. Mol. Cell Biol. 8, 258 (2007). [5] Y. Shibata, J. Hu, M. M. Kozlov, and T. A. Rapoport, Annu. Rev. Cell Dev. Biol. 25, 329 (2009). [6] W. Helfrich, Z. Naturforsch. C 28, 693 (1973). 6 +1/2+1/2−1/2(c)(a)(b)(d)Annihilation of+1/2 and −1/2pair of−1/2+1/2 pair [7] J. B. Fournier, Phys. Rev. Lett. 76, 4436 (1996). [8] J.-B. Fournier and P. Galatola, Brazilian J. Phys. 28, 329 (1998). [9] V. Kralj-Iglic et al., J. Stat. Phys. 125, 727 (2006). [10] G. S. Smith, E. B. Sirota, C. R. Safinya, and N. A. Clark, Phys. Rev. Lett. 60, 813 (1988). [11] A. Tardieu, V. Luzzati, and F. C. Reman, J. Mol. Bio. 75, 711 (1973). [12] Y. Yin, A. Arkhipov, and K. Schulten, Structure 17, 882 (2009). [13] J. Park, T. C. Lubensky, and F. C. Mackintosh, Europhys. Lett. 20, 279 (1992). [14] F. C. MacKintosh and T. C. Lubensky, Phys. Rev. Lett. 67, 1169 (1991). [15] W. Helfrich and J. Prost, Phys. Rev. A 38, 3065 (1988). [16] P. Nelson and T. Powers, Phys. Rev. Lett. 69, 3409 (1992). [17] J. V. Selinger and J. M. Schnur, Phys. Rev. Lett. 71, 4091 (1993). [18] J. M. Schnur, Science 262, 1669 (1993). [19] G. S. Ayton, P. D. Blood, and G. A. Voth, Biophysical Journal 92, 3595 (2007). [20] G. S. Ayton et al., Biophys. J. 97, 1616 (2009). [21] H. Shin, M. Bowick, and X. Xing, Phys. Rev. Lett. 101, 037802 (2008). [22] D. R. Nelson, Nano Lett. 2, 1125 (2002). [23] N. Ramakrishnan, P. B. S. Kumar, and J. H. Ipsen, Phys. Rev. E 81, 41922 (2010). [24] See supplementary materials at http://www.rsc.org/suppdata/sm/c2/c2sm07384f/c2sm07384f.pdf [25] P. A. Lebwohl and G. Lasher, Phys. Rev. A 6, 426 (1972). [26] G. Gompper and D. M. Kroll, Phys. Rev. Lett. 73, 2139 (1994). [27] T. C. Lubensky and J. Prost, J. Phys. II France 2, 371 (1992). [28] M. Bowick and L. Giomi, Adv. Phys. 58, 449 (2009). [29] V. Vitelli and D. R. Nelson, Phys. Rev. E 74, 21711 (2006). [30] Such a quench can be achieved by a change in temperature across the isotropic-nematic transition, wherein the mean bending modulus remains almost the same but a sharp change in the nematic stiffness. [31] S. M. Sweitzer and J. E. Hinshaw, Cell 93, 1021 (1998). 7
1208.3950
1
1208
2012-08-20T09:13:00
Mesoscopic Model for Free Energy Landscape Analysis of DNA sequences
[ "physics.bio-ph", "q-bio.BM" ]
A mesoscopic model which allows us to identify and quantify the strength of binding sites in DNA sequences is proposed. The model is based on the Peyrard-Bishop-Dauxois model for the DNA chain coupled to a Brownian particle which explores the sequence interacting more importantly with open base pairs of the DNA chain. We apply the model to promoter sequences of different organisms. The free energy landscape obtained for these promoters shows a complex structure that is strongly connected to their biological behavior. The analysis method used is able to quantify free energy differences of sites within genome sequences.
physics.bio-ph
physics
Mesoscopic Model for Free Energy Landscape Analysis of DNA sequences. R. Tapia-Rojo,1, 2 D. Prada-Gracia,3 J. J. Mazo,1, 4 and F. Falo1, 2 1Dpto. de F´ısica de la Materia Condensada, Universidad de Zaragoza. 50009 Zaragoza, Spain 2Instituto de Biocomputaci´on y F´ısica de Sistemas Complejos, Universidad de Zaragoza. 50009 Zaragoza, Spain 3Freiburg Institute for Advanced Studies (FRIAS), University of Freiburg, 79104 Freiburg, Germany 4Instituto de Ciencia de Materiales de Arag´on, C.S.I.C.-Universidad de Zaragoza. 50009 Zaragoza, Spain. A mesoscopic model which allows us to identify and quantify the strength of binding sites in DNA sequences is proposed. The model is based on the Peyrard-Bishop-Dauxois model for the DNA chain coupled to a Brownian particle which explores the sequence interacting more importantly with open base pairs of the DNA chain. We apply the model to promoter sequences of different organisms. The free energy landscape obtained for these promoters shows a complex structure that is strongly connected to their biological behavior. The analysis method used is able to quantify free energy differences of sites within genome sequences. PACS numbers: 87.15.A-, 87.15.H-, 87.14.gk I. INTRODUCTION The study of biomolecules at a mesoscopic level tries to identify the main degrees of freedom of the system and understand its behavior in terms of the dynamical and statistical-mechanics properties of the model. At this level, the concept of a free-energy landscape (FEL) rep- resents a paradigm for the comprehension of several bio- logical complex problems such as protein folding, protein structure, and biomolecular interaction [1]. A FEL gives the change in the free energy of a system when the dif- ferent degrees of freedom change. The description given by the topology of the FEL permits to connect struc- ture, dynamics, and thermodynamics in many different systems ranging from atomic clusters to biomolecules or soft matter systems [2]. Here we address our attention to the characterization of the FEL of DNA sequences and, in particular, those sites with regulatory and transcriptional relevance. Re- cently, a great attention has been devoted to the mecha- nism whereby proteins bind to specific sites on DNA [3]. The quantification, grounded in a physical basis, of the strength of these binding sites is an open problem. In this paper, we propose a model which allows us to calculate free energy differences between specific and nonspecific binding sites. Even more, we are able to build, from the trajectories of the model, a representation of the free energy landscape. The model is inspired in the protein search of the bind- ing sites in a DNA chain. This search is a combination of three-dimensional (3D) jumps between separated re- gions of DNA and one-dimensional (1D) diffusion along the chain. It has been stated that most of the search time is spent in the 1D diffusion process since the time jumps in three dimensions is negligible [4 -- 6]. Thus, the restriction to a 1D search is a good starting point for our model. On the other hand, it has also been conjectured that the dynamics of the DNA chain plays an important role in the recognition of binding sites by the regulatory factors or the transcription protein [7, 8]. Thus, tran- scription processes, for instance, would be induced by the binding of RNA-polymerase to openings (bubbles) in the DNA chain. This idea is supported by compu- tational approaches to the DNA dynamics [9], and ex- perimental evidences [10]. Our model follows this idea and considers the interaction of a test particle, which ex- plores the DNA chain, coupled to the bubbles. In order to be physically and biologically relevant, such a descrip- tion should provide useful qualitative and quantitative information about the process. A similar strategy has been used to characterize complex networks and identify regions of special relevance (communities). In such an approach, a "fictitious" Brownian particle goes over the graph [11] and its dynamics reveals "thermodynamics" and structural quantities of the topology [12]. In this paper, we propose a mesoscopic model for DNA- particle interaction. In our picture, the test particle un- dergoes a 1D Brownian motion in interaction with a clas- sical field, the DNA chain itself, whose dynamics is also affected by the presence of the particle. The test particle interacts more strongly with open base-pairs of the DNA chain. In this way "softer" regions of the DNA sequence are more likely to be visited by the particle, which will help also in stabilizing the bubbles. This interaction is not sequence-dependent, as the DNA base-pair dynam- ics already depends on the sequence. Thus, this model could also represent the interaction of a real protein as RNA-polymerase, with the DNA bubbles. Particle and chain are described at the same level of complexity. We use the Peyrard-Bishop-Dauxois (PBD) [13, 14] model to perform the dynamics of the chain. This model was proposed initially for the study of DNA ther- mal denaturation and incorporates the formation and dy- namics of bubbles in a natural way. The PBD model can take into account the sequence information through its parameters. Our model incorporates three basic ingredients of the physics of the system: DNA sequence, bubble dynam- 2 Vint(Xp,{yi}) = − B√ πσ2 (cid:88) i tanh(γyi) e−(Xp−ia)2/σ2 (3) where B sets the interaction amplitude, γ the range of interaction with the base separation and σ the spatial range of interaction on the DNA chain. The functional form for the interaction has been chosen to be linear at low yi and to saturate at large yi in order to avoid that the chain opens indefinitely. Note that with this term the particle is trying to open the chain in a length range of σ and get self-trapped. Although possible, no sequence de- pendence is included in this term since we are interested in giving a general picture of the FEL. We still have to fix the parameters for this interaction term. For the particle damping and mass, we take ηp = 1014s−1 and mP = 7000Da. These values are of the order of magnitude of proteins which bind DNA [18, 19]. The intensity of the interaction chosen is B = 0.52eV . This value provides local interactions of the order of the Morse potential dissociation energy at each base pair. The parameter γ = 0.8A−1 saturates the interaction at y = 1.25A, a typical value for open base pairs. We take a = 1 for the longitudinal separation between base pairs, in arbitrary units, and consider σ = 3. This provides an interaction range of around 5 or 6 base pairs (bp). It is interesting to note that this value has been chosen in order to observe states with bubbles of 10 − 20 bp, which is an adequate width for the processes we take into account here [20]. Once we have fixed the model parameters, we derive the Langevin equations for both the chain bases and the particle. For the chain, we get: m ∂2yn ∂t2 + mη ∂yn ∂t = − ∂ [W (yn, yn+1 + W (yn−1, ym)] ∂yn − ∂V ∂yn − ∂Vint ∂yn + ξn(t), (4) where η stands for the damping and ξn for the thermal noise, so (cid:104)ξn(t)(cid:105) = 0 and (cid:104)ξn(t)ξk(t(cid:48))(cid:105) = 2mηkBT δnkδ(t− t(cid:48)) hold. The Langevin equation for the particle is: mp ∂2Xp ∂t2 + mpηp ∂Xp ∂t = − ∂Vint ∂Xp + ξp(t). (5) where, ηp stands for the particle damping and ξp for the thermal noise. The fluctuation-dissipation relation reads as (cid:104)ξp(t)(cid:105) = 0 and (cid:104)ξp(t)ξp(t(cid:48))(cid:105) = 2mpηpkBT δ(t − t(cid:48)). The equations were numerically integrated using the stochastic Runge-Kutta algorithm [21]. The integration of the Langevin equations of motion provides trajectories of the particle and the DNA chain. Each DNA sequence was simulated in five different realizations for 40µs, us- ing 10 fs time steps and a 1µs preheating time. Since it has been reported that 1D diffusion periods cover a time of the order of milliseconds [18, 19], the simula- tion time used is reasonable for the problem considered FIG. 1. (Color online) Schematic illustration of the DNA- particle interaction model. The unidimensional chain (solid circles) is represented by the base pair opening coordinate (yi) while the probe particle (shaded ellipse) is a diffusing particle along the DNA chain (Xp). ics, and 1D particle diffusion. The analysis of the DNA- particle complex allows us not only to identify possible binding sites, but also to describe the whole structure of the free-energy landscape and determine free-energy dif- ferences between different representative states. We show the validity and usefulness of our approach by studying the FEL of three promoter sequences. In each case, the FEL topology gives insight into the biological behavior of the system. II. MODEL We describe the DNA chain by a modified Peyrard- Bishop-Dauxois model[13 -- 15]. There, the complexity of the DNA molecule is reduced to the study of the dy- namics of N base pairs described by the variables yn defining the distance between the bases. In this frame- In work HDN A = (cid:80)N 2m + V (yn) + W (yn, yn−1) (cid:104) p2 (cid:105) n=1 . n our model V (y) = D(e−αy − 1)2 + Ge−(y−y0)2/b, (1) where the first term accounts for the hydrogen bond in- teraction and the second one for interactions with the solvent [15, 16]. It was shown in [15] that the inclusion of this barrier modifies drastically the duration and sta- bility of the bubbles. The potential W (yn, yn−1) describes the stacking in- teraction between the base pairs along the DNA strand, (cid:16) 1 + ρe−δ(yn+yn−1)(cid:17) K 1 2 W (yn, yn−1) = (yn − yn−1)2. (2) In order to study different DNA sequences, the PBD model includes sequence dependent Morse parameters: Dn, αn. Regarding the DNA chain, we use the set of parameters [17] considered in [15]. The particle is represented by a Brownian particle (see Fig. 1) moving in a one-dimensional space with coordi- nate Xp and interacting with the DNA chain through a phenomenological potential which depends on Xp and the set of coordinates {yi}N p/2mp+Vint(Xp,{yi}) with i=1: HP = p2 3 constructed by discretizing the conformational space ex- plored by the dynamical system and considering the hops between the different configurations as dictated by the MD simulation. In this way, the nodes of a CMN are the subsets of configurations defined by the conforma- tional space discretization, and the links between nodes account for the observed transitions between them. The information of the stochastic trajectory allows us to as- sign probabilities for the occupation of a node (Pi) and for the transitions between two different configurations (Pij). Defined as above, a CMN is thus a weighted and directed graph. It should be stressed that the informa- tion contained in the CMN is much richer that one given by equilibrium statistical mechanics since it includes the dynamics of the system encoded in the probability tran- sitions, Pij. In our case, we start from the reduced trajectory for the DNA (obtained using five principal components) and the trajectory of the test particle. We discretize the to- tal coordinate space in 20 bins of equal volume for the reduced trajectory and N bins (the DNA base pairs) for the particle. This will constitute the microstate space of the CMN, each node with occupancy probability Pi obtained from the reduced trajectory. Once the CMN has been built, we split it into basins of attraction, i.e., regions in which the probability fluxes (Pij) converge to a common state (attractor) of the network. This task is usually hard, since algorithms scale as power law of the system size. In this case we have applied the stochastic steepest descent algorithm developed in [23], which scale as N log N . In this decomposition, a basin corresponds to a coarse-grained state (of connected nodes) of the CMN. In next section, we represent each basin by its attracting node. Once these basins have been defined we can represent the FEL by a hierarchical tree diagram (dendrogram) [23], built according to the weights and links among the basins. This representation is similar to the "disconnec- tivity graph" scheme used in other context [1, 2, 27]. First, an "adimensional free energy" is assigned to each node i given by Fi/kT = log(Pw)− log(Pi) where w rep- resents the weightiest node. Using this magnitude as con- trol parameter, we slowly increase it step by step from its zero initial value. At each step of this process, we obtain a network composed of those nodes with free en- ergy lower than the current threshold value. As the free- energy threshold increases, new nodes emerge together with their links. These new nodes may be attached to any of the nodes already present in the network or they can emerge as a disconnected component. At a certain value of F/kT , some components of the network become connected by the links of a new node incorporated at this step. Initially we have a set of disconnected vertical lines (corresponding to basins) which become linked once the control parameter has overcome the barriers between them i. e. when the free energy of the saddle nodes is reached. Then we draw a horizontal line linking these two basins. Obviously, for large threshold all the net- FIG. 2. (Color online) Trajectory of the particle shown in red line (light gray) with the DNA chain for a collagen sequence (see results section). Black points stands for open bp (yAT > 1.A or yCG > 1.5A) and white for closed ones. here. The simulation temperature is T = 290K and the boundary conditions for the protein are periodic, while for the chain we consider the hard boundary conditions discussed in [15]. An example of such a trajectory is given in Fig. 2. It is observed that the particle moves in a "sea" of open bubbles which clearly shows the soft domains of the genome structure. The particle eventu- ally jumps between these domains opening large stable bubbles. As seen in the figure the dynamics of the bases is strongly affected by the presence of the particle. III. ANALYSIS To extract useful information of such trajectories, and due to the large dimensionality of the system, we ap- ply the principal component analysis (PCA) [22] to the chain trajectory. It has been proved [15] that the first few eigenvectors unveil the softest regions of the DNA chain and hence the possible binding sites for our parti- cle. Even more, PCA reduces the large number of degrees of freedom of the system to just a few, by projecting the coordinates of the system into the first few eigenspaces (reduced trajectories). For each of the sequences consid- ered, we restrict ourselves to the first five eigenspaces. This subspace accounts for 75% of the total fluctuations of the chain dynamics. To obtain the FEL properties of the system we make use of the map of trajectories to a conformational Markov network (CMN) [17,20]. The CMN has been proven to be a useful representation of large stochastic trajec- tories [24 -- 26]. This coarse grained picture is usually 4 FIG. 3. (Color online) Free energy dendrograms (top) for each of the three sequences together with selected states (bottom). The dendrograms represented are a detail of the whole structure showing low energy basins only. The construction of the dendrogram has been performed considering the representative node of each basin. Significant biological states have been searched within the network structure. They correspond to the important basins of the hierarchical free energy organization. work is connected. We can plot this process as a "tree diagram" or dendrogram. Using this representation we can understand qualita- tively and quantitatively the hierarchical organization of the basins and the barriers among them and figure out the behavior of different sequences. This method could be applied to a DNA chain without a particle. However, the inclusion of the particle is es- sential to get the FEL of the system. An analysis of the DNA alone (PBD model) lets us determine the opening probabilities and average position of the chain base pairs, and unveils the softer regions that can indeed be related to sites of biological importance. Nevertheless, the FEL of this model is trivial, as opening events are rare and the chain remains closed for most of the time. The in- clusion of the particle stabilizes the bubbles (as can be observed in Fig. 2) and allows us to go further in terms of predictions. We are able to define relevant states in a precise and systematic way (basins), to predict possible binding sites, and to extract the thermodynamical mag- nitudes related with them, thus characterizing these sites in terms of biological importance. IV. RESULTS To illustrate the method and validate our model, we analyze three different promoter sequences. Promoters are DNA regions in which regulation and initiation of transcription of a gene occurs. Two of them correspond to the so called strong promoters, while the one left is a weak promoter [28]. Strong promoters show a high level of expression in mRNA and usually their sequences are close to the consensus sequence. The strong promoters studied here are the P5 virus promoter, given by the 69 bp sequence shown in [7] and the human collagen type I α2 chain, given by the 80 bp sequence shown in [9]. Finally, the weak promoter is the lac operon regulatory region, whose 129 bp sequence has been taken from [8]. In Fig. 3 (top) we show a detail of the free energy dendrograms for each of the mentioned sequences. The basin structure consists of a big set of low occupied (high energy) basins, and a small set which gathers almost the whole trajectory (see below). This small set of basins is the one shown in Fig. 3. In the botton of Fig. 3 some remarkable states for each of the three promoters are highlighted. The method iden- tifies states with a biological meaning as they correspond to the most important basins. The most significant sites we are dealing with are the transcription starting site (TSS) and the TATA box, although additional promoter sequences can be found depending on the genome. The RNA polymerase binds to the TSS, starting the tran- scription into mRNA. Promoter sequences are usually labeled from the TSS (+1). The TATA-box is found ap- proximately 35 base pairs upstream from the TSS [28]. For the collagen chain, three states have been high- lighted. State A identifies the TSS, showing a bubble in this region with the particle placed just there. States B and C are linked to excitations of other important sites such as the TATA box (state B), see [9]. In the same way, we have found a basin related to the TSS in the case of the P5 chain (state C) and the lac operon (state C), together with other regulatory sites. The arrangement of these basins in the free energy den- drogram informs about the relative free energy between the states and the relation between them. For example, the collagen dendrogram contains three main branches, each one related to each of the three states shown. The P5 promoter shows an analogous structure: two main branches and another one divided into two states (B and C) which are kinetically close. The remaining states of each branch correspond to states similar to that shown, with only slight variations in the chain conformation or in the particle position. The lac promoter shows a clearly different behavior compared with the two strong promoters. From a qual- itative point of view, the arrangement of basins differs from the P5 sequence or the collagen one. To visual- ize quantitatively the difference, the basin occupancy is plotted in Fig. 4. We show the weight of each basin (red bars) for the three sequences together with the ac- cumulated weight (blue line). It is remarkable, in the case of the collagen sequence, that a few basins (25 out of 1661) accumulate almost the whole weight of the net- work (over the 99%). The results for the P5 promoter are completely analogous, a few basins account for most information of the dynamics. These basins are the ones shown in the dendrogram of Fig. 3. When we inspect the bottom graph in Fig. 4 we see a completely different tendency. In the lac network, the distribution of weight among the basins is more uniform. For the collagen and P5 networks we can define a threshold from which the individual contribution to the total weight is negligible. The trajectory is concentrated in around 25 basins and the remainder of the network can be seen as a "background". This "background" is limited to those basins with a weight below 10−3 (horizontal lines in Fig 4). Following this criterion, we can distinguish be- tween specific and nonspecific states. Those basins above the threshold (vertical lines in Fig. 4) may be defined as specific states (with a clear biological function) while those below the threshold may be defined as non-specific states. For the collagen sequence, 25 "specific" basins appear, covering 99.41% of the total trajectory. The P5 and lac sequences show respectively 23 and 88 "specific" basins, which gather 99.38% and 96.91% respectively of the total network weight. Using these definitions, we are able to calculate the relative free energy between states. These magnitudes reveal the "strength" of the different sites in each promoter. It has been reported that specific bind- ing proteins show a greater affinity for strong promoters than for weak ones [29]. To quantify these differences we calculate thermodynamical properties of the most im- portant basins. Once we have divided the network into the different basins of attraction, several statistical mag- nitudes can be defined from them. The weight of the basin is defined as the sum of that of the nodes belong- i Pi with i ∈ α. In the same way the entropy of each basin i Pi log Pi with i ∈ α. Attending to the previous definition of the nonspecific basin, whose thermodynamical magnitudes can be com- puted as explained, we can calculate the free energy of each basin with respect to the nonspecific state. If Pβ is the weight of the non-specific basin, then the free en- ergy difference between a basin α and the macrostate ing to the basin, i. e. for a basin α we have Pα =(cid:80) can be defined as Sα/kB = −(cid:80) 5 FIG. 4. (Color online) Basin occupancy (bars) together with accumulated weight (solid line) for the collagen, P5 and lac sequences. The horizontal line shows the weight threshold be- tween specific and nonspecific states. The vertical line estab- lishes such a frontier in terms of basins. Note the logarithmic y axis. β is ∆Fα/kBT = − log(Pα/Pβ). Table I shows signifi- cant differences between strong and weak promoters. On the one hand we observe that both the total weight and entropy of the non-specific states in the weak promoter exceed by almost an order of magnitude the ones shown for the strong promoters. On the other hand, we can see that the specific states show much higher free energy dif- ferences with respect to the nonspecific states in the case of the strong promoters than the ones shown for the lac sequence. Thus, the analysis presented here opens the way to a systematic study of promoter character within the framework of a mesoscopic model. Promoter State Collagen P5 lac A (TSS) B(TATA) C NS A(TATA) B C (TSS) D NS A(TATA) B C (TSS) D NS Pe 0.169 0.157 0.086 0.006 0.135 0.107 0.086 0.059 0.006 0.115 0.095 0.090 0.038 0.031 S/kB 1.365 1.380 0.652 0.085 1.051 0.913 0.684 0.494 0.027 0.970 0.891 0.775 0.373 0.390 −∆F/kT 3.305 3.232 2.519 0.000 3.130 2.898 2.681 2.301 0.000 1.311 1.120 1.066 0.204 0.000 TABLE I. Statistical (occupation probabilities) and thermo- dynamical (entropy and free energy differences) magnitudes calculated for the chosen states of Fig. 3 and the nonspecific states (NS), according to the criteria shown in Fig. 4. In addition to the three promoter sequences of real bi- ological systems, we have analyzed a random sequence in order to prove the validity of our model. The ran- dom sequence has been obtained by taking the P5 pro- moter sequence and shuffling its base pairs, so that the obtained sequence contains the same base pairs but in random positions. This sequence should contain no ge- netic information at all, and this fact must be reflected in our analysis. When analyzing the random sequence with our method, we can see huge differences compared with the P5 promoter, as we would expect (see Fig. 5). First the structure of the network is completely different. As there are no prominent states in the sequence, the number of basins is huge (8388 compared with the 529 in the P5 promoter). The first two eigenvectors are representative of a homogenous lattice without localized states. The distribution of weights is also clearly different as now the "background" basins keep 6% of the total network weight, an even higher value than that of the background basins gathered in the weak promoter. The dendrogram also shows a much more distributed structure where, even though some nodes appear to fall to privileged positions, their relevance within the whole network structure is far from being comparable to that shown in networks from biological promoters. All these facts validate our model, as we can clearly distinguish between a sequence with binding sites, and thus with biological information, and one with none, even though their chemical composition is the same. V. CONCLUSIONS In this paper we have proposed and analyzed a meso- scopic model for the characterization of binding sites on 6 FIG. 5. (Color online) Analysis of a random sequence. Top: Probability of aperture along with the first two PCA eigen- vectors. Middle: Free energy dendrogram. Bottom: Basin oc- cupancy (bars) together with accumulated weight (solid line) DNA promoter sequences. The model is based on the 1D diffusion of an extended probe particle along the DNA chain. The particle is coupled to the opening states of the chain (bubbles). In its dynamics, it visits the main sites of the sequences, with dwelling times covering a high percentage of the trajectory. Such behavior has allowed us to perform a deep analysis of the FEL which reveals the structure of the complex phase space. The analyzed promoter sequences have been chosen to include genomes from organisms of different domains (virus, bacteria and eukaryote) and different strengths of expression. The model and the analysis used are able to capture the main biological details of the sequences. Our model gives energy differences between specific and nonspecific sites of the promoter. Our results are in good relative agreement with some data in the literature (see for instance [29]): they account for energy ratios be- tween weak and strong promoters. This fact would also make possible the study of sequences in which several TSSs are involved, showing the relative strength between them. We think that our results show the power of coarse- grained or phenomenological mesoscopic models to qual- itatively and quantitatively analyze complex biological systems, in particular the problem of protein-DNA reg- ulatory and transcriptional interactions. Protein-DNA interaction is a fundamental problem which has been the object of a very intense research from many differ- ent points of view in the past years [3, 5]. Our system 7 can be seen as the searching problem of a universal pro- tein on a given DNA sequence, providing an approach for the study of specific protein-DNA interactions at the mesoscopic level, where different protein will interact in different ways with DNA molecules. ACKNOWLEDGMENTS We thank Mar´ıa F. Fillat for helpful discussions on promoter biology. The work is supported by the Spanish Projects No. FIS2008-01240 and No. FIS2011-25167, cofinanced by Fondo Europeo de Desarrollo Regional (FEDER) funds. [1] D. J. Wales Current Opinion in Structural Biology 20 310 (2010); K. Klenin, B. Strodel B, D. J. Wales , W. Wenzel Biochimica et Biophysica Acta 1814 9771000 (2011). E 47 684 (1993). [14] M. Peyrard, Nonlinearity 17 R1-R40 (2004). [15] R. Tapia-Rojo, J. J. Mazo and F. Falo, Phys. Rev E 82 [2] D. J. Wales, Phil. Trans. R. Soc. A 370, 28772899, 031916 (2010). (2012). [3] O. Benichou et al., Phys. Rev. Lett. 103, 138102 (2009); V. Dahirel et al. Phys. Rev. Lett. 102, 228101 (2009); A. Tafvizi , L. A. Mirny, and A. M. van Oijen, ChemPhysChem 12, 1481 (2011). [16] G. Weber, Europhysics Letters 73 806-811 (2006). [17] DAT = 0.052eV, αAT = 4A−1; DCG = 1.5DAT , αCG = 1.5αAT . Gx = 3Dx, y0x = 2/αx, bx = 0.5/α2 x. K = 0.03eVA2, ρ = 3, and δ = 0.8A−1. m = 300Da, η = 5ps−1. [4] K. Sneppen K. and G. Zocchi, Physics in Molecular Bi- [18] M. C. DeSantis, J-L Li, Y. M. Wang, Phys. Rev. E 83 ology (Cambridge University Press, 2005). 021907 (2011). [5] P. H. von Hippel and O. G. Berg, J. Biol. Chem. 264, [19] Z. Wunderlich, L. A. Mirny, Nucl. Acids. Res.86 3570 675-678 (1989) . (2008). [6] S. E. Halford and J. F. Marko, Nucleic Acid Reseach 32 [20] M. Sheinman, O. Benichou, Y. Kafri and R. Voituriez. 3040-3052 (2004). Rep. Prog. Phys. 75 026601 (2012). [7] G. Kalosakas, K. O. Rasmussen, A. R. Bishop, C. H. [21] H. S. Greenside and E. Helfand The Bell System Techni- Choi, A. Usheva, Eur. Phys. Lett. 68 127 (2004). cal Journal 60 1927 (1981). [8] A. Apostolaki and G. Kalosakas, Phys. Biol. 8 026006 [22] I. T. Jolliffe, Principal Components Analysis 2ed (2011). [9] B. S. Alexandrov, V. Gelev, S. W. Yoo, A. R. Bishop, K. O. Rasmussen and A. Usheva PLoS Comput. Biol. 5 e1000313 (2009). [10] C. H. Choi et al., Nucl. Acids. Res. 32, 1584-1590 (2004); B. S. Alexandrov et al., Nucl. Acids. Res. 38, 17901795 (2010); S. Cuesta-Lopez et al. Nucl. Acids Res. 39 52765283 (2011). [11] M. Rosvall and C. T. Bergstrom Proc. Natl. Acad. Sci. U.S.A. 105 1118-1123 (2008). [12] J. G´omez-Gardenes and V. Latora Phys. Rev. E 78 065102 (R) (2008); J-C Delvenne and A-S Libert Phys. Rev. E 83 046117 (2011). [13] T. Dauxois , M. Peyrard, and A. R. Bishop, Phys. Rev. (Springer-Verlag, New York, 2002). [23] D. Prada-Gracia, J. G´omez-Gardenes, P. Echenique, F. Falo, PLoS Comput. Biol. 5 e1000415 (2009). [24] F. Rao, A. Caflisch. J. Mol. Biol. 342 299-306 (2004). [25] A. Caflisch Curr. Opin. Struct. Biol. 16 71-78 (2006). [26] D. Gfeller, P. De Los Rios, A. Caflisch, F. Rao Proc. Natl. Acad. Sci. U.S.A. 104 1817-1822 (2007). [27] S. V. Krivov and M. Karplus, Proc. Natl. Acad. Sci. U.S.A. 101, 14766 (2004); S. Auer et al. Phys. Rev. Lett. 99 , 178104 (2007). [28] R. Schleif, Genetics and Molecular Biology (Addison- Wesley Publishing Company, 1993). [29] L. Bintu, N. E. Buchler, H. G. Garcia, U. Gerland, T. Hwa, J. Kondev and R. Phillips. Curr. Opin. Genetics Dev 15, 116-124 (2005).
1103.1971
1
1103
2011-03-10T09:24:33
Relative velocity of sliding of microtubules by the action of Kinesin-5
[ "physics.bio-ph", "cond-mat.stat-mech", "q-bio.SC" ]
Kinesin-5, also known as Eg5 in vertebrates is a processive motor with 4 heads, which moves on filamentous tracks called microtubules. The basic function of Kinesin-5 is to slide apart two anti-parallel microtubules by simultaneously walking on both the microtubules. We develop an analytical expression for the steady-state relative velocity of this sliding in terms of the rates of attachments and detachments of motor heads with the ATPase sites on the microtubules. We first analyse the motion of one pair of motor heads on one microtubule and then couple it to the motion of the other pair of motor heads of the same motor on the second microtubule to get the relative velocity of sliding.
physics.bio-ph
physics
Relative velocity of sliding of microtubules by the action of Kinesin-5 Sthitadhi Roy1 1Department of Physics, Indian Institute of Technology,Kanpur 208016, India. Kinesin-5, also known as Eg5 in vertebrates is a processive motor with 4 heads, which moves on filamentous tracks called microtubules. The basic function of Kinesin-5 is to slide apart two anti-parallel microtubules by simultaneously walking on both the microtubules [7]. We develop an analytical expression for the steady-state relative velocity of this sliding in terms of the rates of attachments and detachments of motor heads with the ATPase sites on the microtubules. We first analyse the motion of one pair of motor heads on one microtubule and then couple it to the motion of the other pair of motor heads of the same motor on the second microtubule to get the relative velocity of sliding. by another hinge. These two hinges are then linked by a flexible connector. We also assume that the two heads on each side are free to rotate about their respective hinge. The two heads can also pass each other freely. However there is one restriction on their relative displacement. The maximum separation between two heads on one side can be equal to the separation between two discrete bind- ing sites on the microtubule. This separation, which is a constant, is denoted by d. We assume that the micro- tubules are aligned anti-parallel because Kinesin-5 has a much stronger preferences for cross-linking anti-parallel overlapped microtubules than parallel ones [2]. The + end directed stepping activity of Kinesin-5 [4] slides mi- crotubules away from each other. FIG.2 shows schemat- ically the model of one pair of heads on a microtubule. I. INTRODUCTION The protein constituents of the cytoskeleton of eu- karyotic cells can be divided broadly into the follow- ing three categories: (i)filamentous proteins, (ii) motor proteins, and (iii)accessory proteins. The three classes of filamentous proteins which form the main scaffold- ing of the cytoskeleton, are: (a)actin, (b)microtubule, and (c)intermediate filaments. The three superfamilies of motor proteins are: (i)myosin superfamily, (ii)kinesin superfamily, and (iii)dynein superfamily. Microtubules serve as tracks for kinesins and dyneins.[1]. The main cellular mission of the molecular motor Kinesin-5 is to cross-link anti-parallel microtubules and to slide them apart, thus playing a critical role during bipolar spindle formation [2]. Kinesin-5 is absolutely es- sential for the dynamic assembly and function of the mi- totic and meiotic spindles [3]. In this work, I have closely followed the paper by Pe- skin and Oster [5]. II. MODEL The molecular motor Kinesin-5 is a four headed motor with two heads on each side. FIG-1 shows a schematic representation of the model we have developed for this motor. In our model, we assume the two heads of the FIG. 2. Schematic diagram showing only one side of the motor with its two heads at discrete adjacent binding sites on a microtubule. The factors included in the model which are responsi- ble for the motor to move forward are: (i) The back head has a higher probability to detach than the front head. (ii) The attached head prefers a geometry in which it leans in the forward direction i.e towards the + end of the microtubule. The first of these factors ensures that the back head detaches and the second one ensures that it reattaches as a front head [5]. When both the heads are attached to the microtubules then the back head is in a more favourable orientation than the front one. There- fore, ATP hydrolysis and subsequent detachment is more FIG. 1. Schematic diagram of Kinesin-5 on microtubules motor on one side to be joined at a hinge. Similarly, the other two heads on the other side are also joined likely at the back head. Once the back head detaches, the front head relieves itself of the strain by leaning forward and swinging the detached back head forward. Hence this is the process by which the motor moves forward. The coordinates of the two heads are taken to be x = x1 and x = x2. The hinge is located halfway betwween them. The coordinate of the hinge, xh is therefore xh = x1 + x2 2 and the restriction dictates x1 − x2 ≤ d Let α = rate constant for attachment of the free head to the track βb = rate constant for detachment of the back head from the track βf = rate constant for detachment of the front head from the track p = probability that the free head binds infront of the bound head ⇒ 1 − p = probability that the free head binds behind the bound head. Each head is also assigned a number which signifies its state. If it is attached to the track it is assigned 1 and if it is detached from the track it is assigned 0. The state 00 corresponds to both the heads of the cor- responding side completely detached from the track and that implies the termination of the walk. So we only con- sider the states 11, 01 and 10. FIG.3 shows the possible transitions and their rates. FIG. 3. Possible transitions in the states of one pair of heads of the motor and the transition rates In order to monitor the motion of one pair of heads on a microtubule and the velocity of this motion, it is simpler and adequate to monitor the motion of the mid- point of the heads [5]. The average velocity with which the mid-point moves on the microtubule can be taken as the average velocity with which the microtubule slides past the motor heads. We define, xm ≡ coordinate of the mid-point, which is identified as follows [5]: ⇒ xm ≡ position of the hinge in the 11 state ⇒ xm ≡ position of he bound head in the 10 ≡ 01 state 2 So when both the heads are attached then the mid- point is halfway between the two ATPase sites. As soon as one head detaches, the mid-point shifts to the location of the bound head i.e one of the two binding sites on either side of the previous location of the mid-point. So from FIG.2 we can see that the midpoint jumps ± d 2 and this is the step size of the motion of the mid-point. A state diagram showing the possible steps of the mid-point and their rates is shown in FIG.4. It is assumed that the binding sites are labelled by half-integer indices and the mid-points between them are, consequently, labelled by integer indices. So when both the heads are attached, the mid-point is at an integer site and when only one head is attached, the mid-point is at a half-integer site. FIG. 4. The steps the mid-point can take and their respective rates So using the idea of Peskin and Oster explanined above, the motion of a pair of heads on one side can be simplified to the motion of the mid-point of the head alone. Similarly, the same thing can be done for the other pair of heads as well. So instead of monitoring the motion of four heads we can monitor the motion of the two mid-points like a two-headed motor. Now these two mid-points are coupled together because they are parts of the same 4-headed Kinsesin. This constraint introduces a restriction on the difference between the coordinates of the mid-points. If xm1 and xm2 are the coordinates of the mid-points then restriction is xm1 − xm2 ≤ d 2 . So this ensures that if the mid-point is at a site i on a mi- crotubule then cross-linking can take place only through sites i and i ± 1 2 on the other microtubule. Similarly, if the mid-point on a microtubule is at a site i + 1 2 , then the possible cross-linking sites on the other microtubule are i + 1 2 , i and i + 1. The orientations allowed are shown in FIG.5. III. CALCULATIONS AND RESULTS Let Pj (t) ≡ probabilty of finding the mid-point at j at time t. From FIG.4 we can write the following master equations: dPj(t) dt = αpPj− + α(1 − p)Pj+ 1 2 1 2 − (βf + βb)Pj (1) Mean position = PxP (x) Xj=−∞ M eanP osition ⇒ = d ∞ [jPj + (j + ⇒ M eanSpeed d = d dt [ ∞ [jPj + (j + Xj=−∞ 3 1 2 1 2 )Pj+ 1 2 ] )Pj+ 1 2 ]] FIG. 5. The possible orientations of cross-links. The squares represent discrete sites where mid-points can be present. The circles represent the mid-points. < v > d = [ [j dPj dt ∞ Xj=−∞ + (j + 1 2 ) dPj+ 1 2 dt ]. If we define < v > as the average speed of the mid- point then using the above relation and equations (1) and (2) we can write − (βf + βb)Pj ] 1 2(cid:19) [βbPj + βf Pj+1 − αPj+ 1 2 ] 2 1 2 ∞ + + α(1 − p)Pj+ 1 j[αpPj− Xj=−∞ Xj=−∞(cid:18)j + j=−∞ Pj and N0 =P∞ = α(cid:18)p − 2(cid:19) N0 +(cid:18) βb + βf 1 2 d (cid:19) M0 j=−∞ Pj+ 1 2 , we get (t) dPj+ 1 dt 2 Now we define = βf Pj+1 + βbPj − αPj+ 1 2 (2) < v > d = . ∞ j=−∞ Pj+ 1 2 ∞ = dM0 dt dPj dt M0 =P∞ j=−∞ Pj and N0 =P∞ Xj=−∞ Xj=−∞ [αpPj− = ∞ 1 2 + α(1 − p)Pj+ 1 2 − (βf + βb)Pj ] Using M0 =P∞ < v > ⇒ dM0 dt = αN0 − (βf + βb)M0 (3) Similarly, dN0 dt = ∞ Xj=−∞ [βf Pj+1 + βbPj − αPj+ 1 2 ] . Plugging in the expressions for M0 and N0 from equa- tions (7) and (8), we get [5] d 1 < v > = α(cid:18)p − ⇒< v > =(cid:18) αd 2(cid:19)(cid:18) βf + βb 2(cid:19) + α + γ(cid:19)(cid:20)γ(cid:18)p − 1 δ 2(cid:21) α + βf + βb(cid:19) +(cid:18) (βb + βf )α 2(α + βf + βb)(cid:19) (9) ⇒ dN0 dt = (βb + βf )M0 − αN0 (4) where, In the steady state dM0 (3) and (4) we get dt = dN0 dt = 0. So from equations γ = βb + βf δ = βb − βf αN0 = (βf + βb)M0 Also from normalisation we have, M0 + N0 = 1 Solving equations (5) and (6), we get M0 = N0 = α α + βf + βb βf + βb α + βf + βb (5) (6) (7) (8) Now we couple the motion of the two mid-points which move on their MT tracks with < v > calculated above. The motion of a mid-point counts only if it is attached to a microtubule and cross-linked to the other microtubule. So the relative velocity due to the motion of a mid-point on a microtubule is equal to < v > multiplied to the probability of a mid-point being found at any particular site on that microtubule, which is then multiplied by the probability of a mid-point being found at all the corre- sponding possible cross-linking sites, and then summed over all the sites of the first microtubule. This prescription can be mathematically expressed as, ∞ using the definitions of equations (12) and (13) 4 dN dt = 2[(βf + βb)M − αN ] (16) ) ) Similarly, [P (1) i (P (2) i + P (2) i+ 1 2 + P (2) i− 1 2 +P (1) i+ 1 2 (P (2) i + P (2) i+ 1 2 + P (2) i+1) + [P (2) i (P (1) i + P (1) i+ 1 2 + P (1) i− 1 2 < V >= < v >1 < v >2 Xi=−∞ ∞ Xi=−∞ +P (2) i+ 1 2 (P (1) i + P (1) i+ 1 2 + P (1) i+1)] (10) dQ dt = 2 Pi dPi dt ∞ Xi=−∞ Here < v >1 and < v >2 are the velocities of the mid- points on the first and second microtubules respectively. Similarly, the superscripts on the probabilities also refer to the the microtubules. But, since there is no functional difference between the two microtubules, these indices can be dropped. Then we have ∞ using equation (1) ∞ = 2 dQ dt using the definitions of equations (12) and (14) + α(1 − p)Pi+ 1 Xi=−∞ [Pi(αpPi− 1 2 2 − (βf + βb)Pi)] < V >= 2 < v > [ (P 2 i + P 2 i+ 1 2 Xi=−∞ We define three new quantities, ∞ + 4PiPi+ 1 2 ] (11) dQ dt = 2[αM + α(1 − p)M − (βf + βb)Q] (17) M = N = Q = ∞ Xi=−∞ Xi=−∞ Xi=−∞ ∞ PiPi+ 1 2 P 2 i+ 1 2 P 2 i (12) (13) (14) In the steady state dN dt = dQ dt = 0, so we have, 2[(βf + βb)M − αN ] = 0 2[αM + α(1 − p)M − (βf + βb)Q] = 0 (18) (19) Also since the variables M ,N and Q capture all the cross- linking probabilities (FIG.5) so we can write, Plugging these expressions in equation (11), < V >= 2 < v > (4M + N + Q) (15) 2M + N + Q = 1 (20) Now, dN dt = 2 ∞ Xi=−∞ Pi+ 1 2 dPi+ 1 2 dt using equation (2) Solving equations (18), (19) and (20) we get, M = α(βf + βb) 2α(βf + βb) + (βf + βb)2 + α2(2 − p) (21) dN dt = 2 ∞ [Pi+ 1 2 Xi=−∞ (βf Pi+1 + βbPi − αPi+ 1 2 )] Using equations (20) and (21), we get the final expression for < V > as, < V > = 2 < v > (1 + 2M ) i.e., < V > = 2(cid:18) αd α + γ(cid:19)(cid:20)γ(cid:18)p − 1 2(cid:19) + δ 2(cid:21)(cid:18)1 + 2α(βf + βb) 2α(βf + βb) + (βf + βb)2 + α2(2 − p)(cid:19) (22) (23) IV. PHYSICAL INTERPRETATIONS From equation (21) it is obvious that M > 0. There- fore, < V >≥ 2 < v > which is also consistent with our intuitive expectation. Since in our model cross-linking is allowed through sites next to the adjacent sites on mi- crotubules, hence the two mid-points move with respect to each other. This factor actually increases the relative velocity above 2 < v >. The microtubules slide with re- spect to the mid-points with veocity < v > and the mid- points move with respect to each other which increases the overall relative velocity. Next we consider a few special cases to establish that our results are consitent with physical intuition. If the probability of the detachment of the back head and the front head is the same and the probability of the free head binding infront and behind the attached head is also the same, then δ =βf − βb = 0 p = 1 2 (24) (25) In that case, using equations (24) and (25) in equations (9) and (23), we get < v >= 0 and < V >= 0 respec- tively, which is expected on physical grounds. If possible orientations for cross-linking are restricted by imposing the condition that cross-linking is allowed only through adjacent sites on two microtubules, then the relative positions of the mid-points with respect to each other do not change. The relative velocity of slid- ing will simply be the sum of the velocities with which each mid-point hops on the microtubule. The velocites are added because of the anti-parallel orientations of the microtubules. Mathematically, in this case M = ∞ Xi=−∞ PiPi+ 1 2 = 0 (26) and hence, using equation (26) in equation (22) we get < V >= 2 < v >, (27) which agrees with the above prediction. V. COMPARISON WITH EARLIER WORKS Kruse et al.[6] calculated the relative velocity of sliding between two filaments due to the action of a two headed motor. In this work the length of the filaments, which is denoted by L is assumed to be finite. It is assumed that cross-linking can take place only through adjacent sites of the two microtubules. A variable ξ is defined, which is the differnce between the coordinates of the minus ends of the microtubules. Thus overlaps can take palce if ξ = 0, ..., 2L − 2. The average velocity of sliding as a function of ξ in units of monomers per unit time is v(ξ) = 2pcl( P1+ξ PL−1 i=1 JiS2+ξ−i i=ξ−L+2 JiS2+ξ−i forξ = 0, ..., L − 2, forξ = L − 1, ..., 2L − 3. (28) where pcl is the probability of cross-linking. When averaged over all possible values of ξ the final expression for average relative velocity is [6] 5 < V >= 1 2L − 1 2L−3 Xξ=0 v(ξ) (29) Now, equation (29) can be used for our system by taking the limit of L → ∞ and replacing v(ξ) by 2 < v >. In that case, < V > = lim L→∞ 1 2L − 1 = lim L→∞ 1 2L − 1 = lim L→∞ 2 < v > 2L − 1 = 2 < v > lim L→∞ < V > = 2 < v > v(ξ) 2 < v > 2L−3 2L−3 Xξ=0 Xξ=0 Xξ=0 2L−3 1 2L − 3 2L − 1 which is consistent with equation (23) because in the spe- cial case cross-linking was allowed only through adjacent sites on two microtubules. Computer simulations as well as analytical studies re- lated to continuum models of motor induced sliding of mi- crotubules and actin filaments have been reported [8, 9]. These studies say that two headed bipolar motors slide anti-parallel filaments with a velocity that is twice that of the free motor velocity [8]. In our case, the motion of Kinesin-5 is reduced to a motion of a two headed motor using the model of mid-points. In our case cross-linking is allowed through adjacent sites on microtubules or sites next to adjacent sites on microtubules. In the continuum limit the distance between two sites is negligible, so it can be said that all the cross-linking that takes place is effec- tively through adjacent sites only, as in the continuum limit the adjacent site and the next to adjacent site can be deemed to be the same. Hence all the cross-linking can be deemed to be through adjacent sites. In such a case as we calculated in equation (27) our realtive veloc- ity of sliding < V > is equal to 2 < v >, where < v > is the velocity with which a mid-point slides a filament. If < v > is taken analogous to the free motor velocity of a two headed motor, then our result in the continuum limit agrees with earlier studies done with continuum models. VI. ACKNOWLEDGEMENTS I sincerely thank Debashish Chowdhury for introduc- ing me to molecular motors, and for useful suggestions. I also thank him for his comments on earlier versions of this manuscript. 6 [1] "Resource Letter PBM-1:Physics of Bio-Molecular Ma- chines", Debashish Chowdhury, Am. J. Phys 77, 583, (2009) [2] "Walking, hopping, diffusing and braking modes of kinesin-5", Kuniyoshi Kaseda, Andrew D.McAnish and Robert A.Cross, Biochem. Soc. Transactions 37, 1045- 1049, (2009) [3] "Kinesin-related proteins required for assembly of the mi- totic spindle", D.M.Roof, P.B.Meluh and M.D.Rose, J. Cell Biol. 118, 95-108, (1992) [4] "Mitotic spindle oraganisation by a plus-end directed mi- crotubule motor", K.E.Sawin, K.LeGuellec, M.Philippe and T.J. Mitchison, Nature 359, 540-543, (1992) [5] "Coordinated hydrolysis explains the mechanical be- haviour of Kinesin", Charles S. Peskin and George Oster, Biophys. J. 68, 202s-211s, (1995) [6] "Growth of finger-like protrusions driven by molecular motors", K. Kruse and K. Sekimoto, Phys. Rev. E 66, 031904, (2002) [7] "The bipolar mitotic kinesin Eg5 moves on micro- tubules that it cross-links", L.C.Kapitein, E.J.Peterman, B.H.Kwok, J.H.Kim, T.M.Kapoor, C.F.Schimdt, Nature 435, 114-118, (2005) [8] "Motor induced sliding of microtubule and actin bundles", Assaf Zemel and Alex Mogilner, Phys. Chem. Chem. Phys. 2009,4821-4833 (2009) [9] "Actively contracting bundles of polar filaments", K.Kruse and F.Julicher, Phys. Rev. Lett. 85:88, 1778-1781, (2000)
1107.0780
2
1107
2011-11-21T19:22:51
Kinetic Monte Carlo and Cellular Particle Dynamics Simulations of Multicellular Systems
[ "physics.bio-ph" ]
Computer modeling of multicellular systems has been a valuable tool for interpreting and guiding in vitro experiments relevant to embryonic morphogenesis, tumor growth, angiogenesis and, lately, structure formation following the printing of cell aggregates as bioink particles. Computer simulations based on Metropolis Monte Carlo (MMC) algorithms were successful in explaining and predicting the resulting stationary structures (corresponding to the lowest adhesion energy state). Here we present two alternatives to the MMC approach for modeling cellular motion and self-assembly: (1) a kinetic Monte Carlo (KMC), and (2) a cellular particle dynamics (CPD) method. Unlike MMC, both KMC and CPD methods are capable of simulating the dynamics of the cellular system in real time. In the KMC approach a transition rate is associated with possible rearrangements of the cellular system, and the corresponding time evolution is expressed in terms of these rates. In the CPD approach cells are modeled as interacting cellular particles (CPs) and the time evolution of the multicellular system is determined by integrating the equations of motion of all CPs. The KMC and CPD methods are tested and compared by simulating two experimentally well known phenomena: (1) cell-sorting within an aggregate formed by two types of cells with different adhesivities, and (2) fusion of two spherical aggregates of living cells.
physics.bio-ph
physics
Kinetic Monte Carlo and Cellular Particle Dynamics Simulations of Multicellular Systems Elijah Flenner,1 Lorant Janosi,1 Bogdan Barz,1 Adrian Neagu,1, 2 Gabor Forgacs,1, 3, 4 and Ioan Kosztin1, ∗ 1Department of Physics and Astronomy, University of Missouri, Columbia, MO 65211 2Department of Biophysics and Medical Informatics, University of Medicine and Pharmacy Timisoara, 300041 Timisoara, Romania 3Department of Biological Sciences, University of Missouri, Columbia, MO 65211 4Department of Biomedical Engineering, University of Missouri, Columbia, MO 65211 (Dated: September 3, 2018) Computer modeling of multicellular systems has been a valuable tool for interpreting and guid- ing in vitro experiments relevant to embryonic morphogenesis, tumor growth, angiogenesis and, lately, structure formation following the printing of cell aggregates as bioink particles. Computer simulations based on Metropolis Monte Carlo (MMC) algorithms were successful in explaining and predicting the resulting stationary structures (corresponding to the lowest adhesion energy state). Here we present two alternatives to the MMC approach for modeling cellular motion and self- assembly: (1) a kinetic Monte Carlo (KMC), and (2) a cellular particle dynamics (CPD) method. Unlike MMC, both KMC and CPD methods are capable of simulating the dynamics of the cellular system in real time. In the KMC approach a transition rate is associated with possible rearrange- ments of the cellular system, and the corresponding time evolution is expressed in terms of these rates. In the CPD approach cells are modeled as interacting cellular particles (CPs) and the time evolution of the multicellular system is determined by integrating the equations of motion of all CPs. The KMC and CPD methods are tested and compared by simulating two experimentally well known phenomena: (1) fusion of two spherical aggregates of living cells, and (2) cell-sorting within an aggregate formed by two types of cells with different adhesivities PACS numbers: 87.17.Aa, 87.17.Rt, 87.85.G-, 87.85.Lf I. INTRODUCTION Understanding how living cells form tissues and organs is a fundamental problem of developmental biology [1, 2], and is also important for the rapidly expanding field of tissue engineering that aims at building functional tis- sue substitutes in vitro [3]. Tissue engineered structures may be used for drug testing and to restore or replace damaged tissues and organs [4]. An emerging tissue engi- neering technique is bioprinting [5 -- 11] via the automated layer-by-layer deposition of multicellular aggregates (the bioink). Subsequent postprinting fusion of the contigu- ous aggregates gives rise to the desired tissue construct. Predicting the result of post-printing tissue formation is a task for theoretical modeling. A guiding principle for most models of cell rearrange- ment in cell aggregates is the differential adhesion hy- pothesis (DAH) proposed by Steinberg [12, 13]. DAH states that structure formation in multicellular systems occurs due to (i) differences in cell-to-cell adhesion of different types of cells and (ii) cell motility. Cells seek positions with the largest number of strong bonds. For example, in a random mixture of two cell types of differ- ent cohesivities the more cohesive cell population sorts out and occupies the central region surrounded by the less cohesive population. ∗ [email protected] By incorporating DAH, Metropolis Monte Carlo (MMC) simulations correctly predict the formation of multicellular structures of minimum energy of adhesion, and identify long-lived, metastable configurations [14]. However, MMC cannot predict the actual time evolution of multicellular systems. Insight into time evolution emerged from experiments designed to verify DAH, which revealed that embryonic tissues behave analogously to highly viscous liquids. The concept of tissue liquidity motivated a quantitative de- scription of embryonic tissues in terms of parameters of continuum hydrodynamics, such as surface tension and viscosity. For example, the (apparent) surface tension γ was measured for several tissue types using a paral- lel plate compression apparatus [15] and the values were found to be consistent with the sorting behavior of these tissues [16]. Here, by using the concept of tissue liquidity and DAH, we formulate two alternatives to the MMC approach for modeling cellular motion and self-assembly: (1) a kinetic Monte Carlo (KMC), and (2) a cellular particle dynam- ics (CPD) method. Both these methods are capable of describing and predicting the real time evolution of the shape of multicellular systems and tissue constructs in certain morphogenetic processes and post-bioprinting structure formation. In the KMC method the configu- ration of the multicellular system is propagated in time through a standard rejection-free kinetic Monte Carlo al- gorithm. This approach should provide a more accurate description of the time evolution of a multicellular system than other grid based methods, such as, the MMC model [6, 14] or the widely used cellular Potts model (CPM). The latter uses a modified Metropolis MC algorithm to update the configuration of the simulated system and postulates that time is proportional to the number of MC steps, which in general is not the case [17]. In the CPD method [17] individual cells are modeled as an en- semble of cellular particles (CPs) that interact via short range contact interactions, characterized by an attractive (adhesive interaction) and a repulsive (excluded volume interaction) component. CPs in a cell are held together by an additional confining potential that mimics the role of the cell membrane. The time evolution of the spatial conformation of the multicellular system is determined directly by recording the trajectories of all CPs by inte- grating their equations of motion. What sets apart CPD from the other similar off-grid particle methods, such as Newman's subcellular element method (SEM) [18 -- 21], is the employed force field (especially the confining poten- tial) and its parametrization that makes the system be- have as a complex viscous liquid. In particular, the CPD model parameters are determined such that the shape of two fusing spherical aggregates in the CPD simulation match as closely as possible the one observed experimen- tally, i.e., two attached spherical caps (see Fig. 1) [22]. To test and compare the KMC and CPD methods, we apply them to simulate the fusion of two spherical aggre- gates and the evolution of cell sorting within an aggre- gate, two morphogenetic processes driving postprinting structure formation. For the theoretical description of the fusion of two identical spherical aggregates we use a simple continuum model introduced by Frenkel [23] and further developed by others working in the field of rhe- ology [24, 25]. It is this theoretical continuum model that provides the link between the time scales of simula- tions and the time scales of experiments. Once this link is established, the KMC and CPD simulations are used to quantitatively predict the time evolution of complex postprinted structures whose description using a contin- uum hydrodynamics approach is impractical. The remainder of the paper is organized as follows. Section II describes the KMC (Sec. II A) and the CPD (Sec. II B) methods, as well as the theoretical aspects of the continuum approach of aggregate fusion (Sec. II C). Section III contains the results and discussion of our KMC and CPD simulations, i.e., fusion of identical spher- ical multicellular aggregates (Sec. III A) and cell sorting (Sec. III B). Conclusions are presented in Sec. IV. II. COMPUTER AND THEORETICAL MODELING A. Kinetic Monte Carlo for Multicellular Systems 2 approaches equilibrium, or is in a metastable state, the Metropolis algorithm rejects most trial moves because the acceptance probability is small. A main feature of the KMC algorithm is that it is "rejection-free". In each step, one calculates the transition rates for all possible changes compatible with the current configuration, and then chooses a new configuration with a probability pro- portional to the rate of the corresponding transition. We designed and implemented a KMC algorithm to simulate the time evolution of a lattice model of multicel- lular systems. Aggregates of cells in cell culture medium are represented on a 3D hexagonal close-packed lattice by associating each site to either a cell or to a similar sized volume element of medium. Thus, the lattice spac- ing is equal to one cell diameter. We assume that each cell interacts with its 12 nearest neighbors (1st and 2nd neighbors considered to be nearest neighbors) located at a distance of one lattice spacing from the given cell. lnter- actions are expressed in terms of works of cohesion and adhesion [27, 28], defined as the work needed to break up the contact between two neighbors of respectively simi- lar or differing types of cells. For example, in case of a multicellular aggregate composed of a single cell type, the work needed to extract a cell from the aggregate (i.e. model tissue) is the work of cohesion, cc, multiplied by the number of the cell's nearest neighbor. The interac- tion between cells and the cell culture medium is set to zero. The movement of cells is described by assigning rates to swapping cells with adjacent cells of different type and/or with medium elements. These elementary moves occur with rates given by k = w0e −Eb/ET , (1) where the factor w0 is the frequency of attempts to cross the energy barrier of height Eb, and ET is the energy of biological fluctuations [29], the analog of the energy of thermal fluctuations, kBT (kB is Boltzmann's con- stant and T is the absolute temperature). It has been argued that ET is a characteristic measure of cell motil- ity: the higher is ET in comparison to the energies of cohesion/adhesion, the higher is the motility of the cell [29]. Due to the complexity of the cytoskeletal machinery responsible for cell movement, there is no unique way to assign a barrier height to the swapping of two cells. Any reasonable choice, however, needs to be consistent with the following set of experimental observations on cell movement in 3D: (1) Relocation of cells in embryonic tissues and in some engineered tissues (such as cell aggregates) occurs according to DAH [13, 30]: cells take advantage of their motility to establish the maximum number of strong bonds with their neighbors. The Kinetic Monte Carlo method (KMC) was pro- posed as an alternative to the MMC method for simu- lating the evolution of Ising models [26]. When a system (2) Anchorage-dependent cells do not spontaneously dissociate from the cell aggregate they are part of [13]. (3) The speed of cell movement in 3D matrices has a particular dependence on the strength of cell- matrix adhesion: cell movement is fastest at an op- timal strength of binding. Too weak or too strong binding hampers cell movement [31, 32]. Consider a binary particle model for a multicellular system formed by two cell types, t = 1, 2 (for a multi- cellular aggregate composed of one cell type, t = 1, sur- rounded by tissue culture medium, t = 2 represents the medium particle). The configurational energy (or total interaction energy), E is expressed as [6] 2 E = γ12N12 + const, 2, and 12 is the energy of adhesion. N12 =(cid:80)N I (cid:80)N II (2) where γ12 = (11 + 22)/2 − 12, with 11 and 22 being the energies of cohesion respectively for cell type 1 and i=1 ni2 = i=1 ni1 is total number of nearest neighbor pairs of different cell types cells, ni2 (ni1) the number of nearest neighbors of cell i of type 1(2), which are of type 2(1) and N I 2 ) the total number of cells of type 1(2), which have at least one (nearest) neighbor of type 2(1). (As the const is irrelevant for the evolution of the system, we set it to zero [6].) 1 (N II 1 Consider two nearest neighbor cells, i and j of different types (without loss of generality we can set i = 1 and j = 2). The system evolves in time towards configurations of decreasing energy E, i.e. for γ12 > 0 (γ12 < 0) N12 decreases (increases). For γ12 > 0 and γ12 < 0 cells respectively phase separate (cell sorting) and mix (cell mixing). Elementary KMC moves consist of swapping two neighbors of different types (swapping cells of same type does not change the energy). The contribution of two such cells, i and j to E is Eij = 1 2 (ni2 + nj1), (3) and E =(cid:80)N I (cid:80)N II 2 1 i=1 j=1 Eij. Furthermore, the larger is Eij the more likely is the KMC move to swap cells i and j. Thus it is reasonable to define the energy barrier Eij b in Eq. (1), for a transition involving the swapping of two cells i and j, as 0≤Eij b = Emax ij − Eij, (4) ij where Emax γ12 > 0, Emax of differing type surrounding cells i and j is maximal. is the maximum possible value of Eij. For is obtained when the number of neighbors ij Now we can formulate the steps of our KMC algo- rithm for simulating the time evolution of multicellu- lar systems: (S1) Set t = 0; (S2) Find all interfacial cells (i.e., cells in contact with cell culture medium or with cells of different type) and compute the rates km, 1 ≤ m ≤ M , corresponding to all possible M transitions involving these cells; (S3) Calculate the cumulative rates: n=1 kn, 1 ≤ m ≤ M ; (S4) Generate a uniform random number u between 0 and 1 and carry out event Km =(cid:80)m 3 "m" for which Km−1 < uKM ≤ Km; (S5) Generate an- other uniform random number u(cid:48) between 0 and 1, and increment the time variable (i.e., t → t + ∆t) by the non-uniform time step ∆t = −K −1 M log(u (cid:48)); (5) (S6) Update all rates kn that may have changed due to the previous transition "m"; (S7) Return to step S2 and repeat the process until the time variable reaches the desired target value. B. Cellular Particle Dynamics Method for Multicellular Systems The cellular particle dynamics (CPD) method is an off-lattice, particle-based computer simulation method that can describe and predict the time evolution of 3D multicellular systems during shapechanging biomechan- ical transformations [17]. Within the CPD formalism cells, regarded as continuous objects with self-adaptive shape, are coarse-grained into a finite number, NCP , of equal volume elements. Each volume element is repre- sented by a point-like cellular particle (CP) situated at its center of mass. CPs interact via short-range contact interactions, characterized by an attractive (adhesive in- teraction) and a repulsive (excluded volume interaction) component. In addition, CPs within a given cell are sub- ject to a confining potential that assures the integrity of the cell. The time evolution of the spatial conformation of the multicellular system is determined directly by cal- culating the trajectories of all CPs (and, therefore, cells) through integration of their overdamped Langevine equa- tions of motion. This minimalist model, when properly parametrized, has the features of a complex viscous liq- uid and it is suitable for describing the time evolution of multicellular aggregates and soft-tissue constructs. For the nth CP in cell α, the equation of motion is µ rαn(t) = −∇αn U + fαn (t), (6) where rαn(t) is the position vector, U is the potential en- ergy function describing the interaction of the CPs , µ is the friction coefficient, fαn (t) is a random force, and the dot denotes time derivative. The random force is mod- eled as a Gaussian white noise with zero mean and vari- ance (cid:104)fi(t)fj(0)(cid:105) = 2Dµ2δ(t)δij, where D is the sort-time self diffusion coefficient of the CPs. The CPD parame- ters D and µ are related to the previously introduced biological fluctuation energy ET by the Einstein relation Dµ = ET . The CP potential energy U has an intra- cellular and an inter-cellular component corresponding to CPs belonging respectively to the same cell and to different cells, i.e., (cid:88) (cid:88) U intra(rαn − rαm) (cid:88) (cid:88) U inter(rαn − rβm), α n=1 m(cid:54)=n U = 1 2 + 1 2 α β(cid:54)=α n,m (7) where αn (βm) labels the cellular particle n (m) in cell α (β). We model the short-range contact inter- and intra- cellular interactions between CPs through (cid:0)r; inter, σinter(cid:1) , U inter(r) = VLJ (8a) U intra(r) = VLJ where (cid:0)r; intra, σintra(cid:1) + (cid:20)(cid:16) σ (cid:17)6(cid:21) (cid:17)12 −(cid:16) σ r r k 2 (r − ξ)2 Θ(r − ξ), (8b) VLJ (r; , σ) = 4 (8c) is the standard Lennard-Jones potential, and Θ(r) is the Heaviside step function. Note that instead of VLJ one could use any other potential that has a repusive core and a short range attractive part. For example, in SEM [18] a Morse potential is used to describe the interaction between two subcellular elements. However, the impor- tant addition in CPD is the second quadratic term in U intra that, for r > ξ, represents an elastic confining po- tential used to maintain the integrity of the cell. This term, characterized by the elastic constant k, guarantees that the CPs within a cell remain confined inside the boundary of the cell. The time evolution of the multi- cellular system within the CPD approach is determined by numerically integrating the equations of motion (6) for all CPs. We have accomplished this by implementing the intra- and inter-cellular interaction forces, Eqs. (7)- (8), and a Langevin dynamics integrator in the freely available massively parallel molecular dynamics package LAMMPS [33]. 1. CPD units and parameters For a multicellular system of a single cell type there are nine CPD model parameters that need to be determined: NCP (the number of CPs per cell), D, µ, σintra, εintra, k, ξ, σinter and εinter. The choice of NCP is determined by the degree of detail we want to describe individual cells. Since we are interested in the time evolution of the shape of an aggregate formed by a large number of cells and not in the detailed description of the surface dynamics of individual cells, in the present work we make the reasonable choice of NCP = 10. Because σ in (8c) determines the size of the interacting CPs, we can set σ ≡ σintra = σinter. The length ξ in (8b) represents the size (diameter) of a cell, which comprises NCP tightly packed CPs of size σ. Thus, one can estimate ξ ≈ σN 1/3 CP . 4 Next, we define convenient CPD (or computer) length, time and energy units according to (cid:96)0 = σ, t0 = σ2 D , E0 = ET = µD. (9) In these units all CPD parameters all pure numbers, and in particular σ = D = µ = 1. We set the confining poten- tial parameters as ξ = 2.5 (∼ 101/3) and k = 5. A larger (smaller) value for k makes the cell more rigid (soft) when subjected to deformations. The chosen value, for which kσ2/2 = 2.5ET , is suitable when cells in the aggregate are exposed only to adhesion and surface tension forces. By choosing intra = 1 (i.e., same as the biological fluc- tuation energy ET ) the dynamics of the CPs inside a cell will have sufficient randomness to produce cell surface fluctuations that play an important role in cell motility [34]. Thus, out of the nine CPD parameters we are left with only one, inter, that needs to be determined such that the time evolution of the shape of the multicellular sys- tem follows as closely as possible the corresponding ex- perimental one. For this purpose we focus on the fusion of two identical spherical aggregates, as described in the next section. Based on extensive CPD simulations we have found that the best agreement with experiment is obtained for 1 < inter < 2, when the system behaves as a viscous liquid. By increasing inter above 2 the fusing cellular aggregates show sign of solidification and their behavior deviate significantly from experiment. The re- sults reported in this paper are for inter = 1. However, these are similar to the ones obtained for any inter < 2. Furthermore, in all our CPD simulations we have used an integration time step ∆t = 10−4, and used a cutoff radius Rc = 2.5 for U inter(r). C. Continuum Description of the Fusion of Two Spherical Cell Aggregates The fusion of two contiguous cell aggregates is driven by surface tension, γ, and resisted by viscosity, η. It is an experimental fact that during the fusion of identical spherical soft tissue aggregates the shape of the system is that of two touching spherical caps (see Fig. 1) [22]. This observation suggests that soft tissues behave like complex viscous liquids whose description requires an a priori un- known hydrodynamic constitutive model. However, the simplicity of the geometry allows us to describe analyt- ically the dynamics of the considered fusion process by employing conservation laws as proposed by Frenkel [23] and Eshelby [35] for the coalescence (sintering) of highly viscous molten drops. The fusing aggregates are modeled as two spherical caps of radius R(θ) with circular contact ("neck ") region of radius r(θ) = R(θ) sin θ (see Fig. 1A). Volume conser- vation requires R(θ) = 22/3(1 + cos θ)−2/3(2 − cos θ)−1/3R0 , (10) 5 Note that according to Eqs. (12) and (14) the dynamics of the fusion process, described by θ(t) as a function of t/τ , is independent of the size (i.e., R0) of the fusing spheres. R0 appears only in the characteristic fusion time τ , Eq. (13). Finally, using Eq. (14), the square of the radius of the circular neck region of the fusing spherical caps can be (15a) expressed as(cid:18) r A(t) = 24/3(cid:16) with R0 (cid:19)2 ≈ A(t)[1 − exp(−t/τ )] , −t/2τ(cid:17)−2/3 −t/2τ(cid:17)−4/3(cid:16) 2 − e 1 + e . (15b) For short times, t (cid:28) τ , Eqs. (15) yield the familiar linear-in-time expression, (r/R0)2 ≈ t/τ , obtained by Frenkel [23] and Eshelby [35]. While this formula has been applied previously to estimate the capillary veloc- ity vc = γ/η = R0/τ of soft tissues [22, 34], we are not aware of any previous study that followed the time evolu- tion of the shape and of [r(t)/R0]2 throughout the fusion process of two spherical tissue aggregates. First, we have determined the dimensionless CPD parameters (i.e., ex- pressed in CPD units; see Sec. II B 1) such that the shape of the fusing aggregates during CPD simulation resemble as close as possible to spherical caps. Second, we have determined the characteristic fusion time τ by fitting the data for [r(t)/R0]2, obtained respectively from experi- ment, CPD and KMC simulations, to Eqs. (15). Finally, the CPD time unit can be calculated as t0 = τexp/τCP D. Once t0 is known, one can predict through CPD simu- lation the time evolution of an arbitrary 3D tissue con- struct built from the same type of cells for which the time calibration was performed through the above method (i.e., fusion of spherical aggregates). Clearly a similar calibration strategy can be used for the KMC method. III. RESULTS AND DISCUSSION To test and compare the KMC and CPD methods described in Sec. II, we have applied them to simulate two important morphogenetic processes: (A) tissue fu- sion (the fusion of two identical spherical multicellular aggregates), and (B) cell-sorting (within a spherical mul- ticellular aggregate formed by two types of cells with dif- ferent adhesivities). FIG. 1. (color online). (A) The shape of two fusing spherical cell aggregates can be quantified by the angle θ(t) and the radius R(t). (B) Throughout the fusion of two indentical spherical cushion tissue aggregates [22] the system has the shape of two connected spherical caps. The numbers indicate the time (in minutes) elapsed from the start of the fusion. with R0 = R(0). Thus, the time evolution of the fu- sion process is parametrized by a single angle θ = θ(t), defined in Fig. 1A, that changes from θ(0) = 0 to θ(∞) = π/2. The rate of the decrease in surface en- Ws = γdS/dt, where the free surface area S = ergy is S(θ) = 4πR2(θ)(1 + cos θ). The equation of motion for θ(t) can be derived by equating Ws with the rate of the energy dissipated by the viscous flow Wη ≈ −4πR3 0ηα2 [23, 25]. Assuming biaxial stretching flow, α = ∂vx ∂x ≈ − 1 R(θ) d dt [R(θ) cos θ]. (11) Inserting Eq. (11) into the energy balance equation Ws = Wη leads to [25] dθ dt = sin θ cos θ(2 − cos θ)1/3 25/3(1 − cos θ)(1 + cos θ)1/3 1 τ = 1 2τ R0 cot θ R(θ) , where the characteristic fusion time τ = ηR0/γ . (12) (13) Equation (12) can be solved numerically for θ = θ(t). However, one can derive a simple and accurate analytical approximation for θ(t) by setting R(θ) ≈ R0 in Eq. (12). Indeed, throughout the fusion process 1 ≤ R(θ)/R0 ≤ 21/3 ≈ 1.26 holds. With this approximation, Eq. (12) can be easily integrated with the result cos θ = exp(−t/2τ ) . (14) A. Fusion of Two Spherical Cell Aggregates As descirbed in Sec. II C, the fusion of two identical spherical aggregates can quantitatively be characterized by the time dependence of the radius, r(t), of their circu- lar contact region. According to Eqs. (15), r(t) obtained from experiment and from KMC and CPD simulations, θxyrRB A can be used to determine the characteristic fusion time τ , Eq. (13). Thus, for a given cell type, by comparing the experimental τ with that obtained from computer simulations one can calibrate the time scale of the cor- responding computer model. Once such a calibration is done, one can make quantitative in silico predictions of the time evolution of various multicellular processes that involve the same cell type [17]. In this section we present KMC and CPD simulation results for the fusion of two identical spherical aggregates. We show that in both cases the computed (r/R0)2 vs t/τ dependence can be reasonably well fitted by Eqs. (15). Then, using experimental results for aggregate fusion [22], the calibration of the KMC and CPD simulation time scales is exemplified for the case of cardiac cushion tissue (CT). Finally, KMC and CPD simulations are used to predict the formation of a toroidal structure by cell ag- gregate fusion, an important structure in the engineering of tubular tissue constructs [8]. 1. KMC simulations The initial radius of the two identical fusing aggregates used in our KMC simulation was R0 = 10 cell diame- ters. Each aggregate contained 5,927 cells, with a cell-cell work of cohesion cc = 0.9. The medium-medium (cell- medium) work of cohesion (adhesion), mm (cm), was considered to be negligibly small. A total of 10 KMC simulations of the same fusion process were carried out, each time using a different seed of the random number generator. Each simulation was run for 105 KMC time steps. Representative snapshots during the KMC fusion sim- ulation are shown in Fig. 2. The corresponding (r/R0)2 vs t/τ dependence is shown in Fig. 3 (dashed curve). Apart from the beginning of the fusion process (i.e., t < τ ) the KMC result appears to match rather well both the theoretical prediction (thick-solid curve), Eqs. (15), and the experimental results corresponding to the fusion of CT aggregates (open-circle) [22]. −1 0 , the fusion time [ob- tained by fitting the KMC simulation results to the the- oretical formula Eqs. (15)] was τ0 = 1.1 × 109. Since the experimental characteristic fusion time for CT ag- gregates τexp ≈ 5h [22], it follows that the KMC time unit (for CT aggregates used in [22]) has the calibrated value t0 = w 0 = τexp/τ0 = 1.6 × 10−5s. −1 In the KMC time unit t0 = w 2. CPD simulations Each of the two spherical aggregates used in the CPD simulation of aggregate fusion contained 2000 cells. The used CPD parameters and integration timestep, in CPD units, are described in Sec. II B 1. The equilibrated ag- gregates were placed within a distance of one σ before starting the fusion simulation. 6 FIG. 2. Time evolution of the fusing aggregates in the KMC (left) and CPD (right) simulations. The snapshots were taken at: (a) t = 0, (b) t = 0.19τ , (c) t = 2.8τ , and (d) t = 5.5τ . The solid-line contours represent the theoretical shapes of the fusing aggregates determined by Eqs. (15). Representative snapshots during the fusion process are shown, and compared with the corresponding KMC sim- ulation results, in Fig. 2. While in both KMC and CPD simulations the profiles of the fusing aggregates for in- termediate stages of the fusion process (Fig. 2b-c) agree quite well, these show noticeable differences with respect to the theoretical prediction, Eqs. (15), shown as solid- line contours in Fig. 2. The (r/R0)2 vs t/τ dependence in the CPD simula- tion is also shown in Fig. 3. The CPD and KMC simula- tion results are similar. Apart from short times (t < τ ) they agree quite well with both the theoretical prediction, Eqs. (15), and the experimental results for CT [22]. The characteristinc CPD fusion time is determined to be τ ≈ 540t0. By equating this with τexp ≈ 5h, one finds that the CPD time unit calibrated for CT aggregates is t0 ≈ 0.6 min. The CPD simulations were preformed on 32 CPUs of a dual core 2.8GHz Intel Xeon EM64T cluster with a performance of around 5 million timesteps/day (which is equivalent to 500t0 and slightly less than 1τ ). CPDKMC(a)(b)(c) (d) 7 FIG. 4. KMC (left) and CPD (right) simulations of toroidal structure formation through the fusion of 10 cell aggregates. Top view of the fusing aggregates at (a) the beginning (t = 0), and (b) the completion of fusion. (c) Cross-section through the median plane of the fused toroidal structure shown in (b). Otherwise identical cells, initially located in adjacent aggregates are colored differently to emphasize the degree of mixing during fusion. by calculating the time dependent mixing parameter dmix(t) = 4 M ∆N L m(t) · ∆N R [∆Nm(t)]2 m(t) . (16) M(cid:88) m=1 m(t) [∆R Here ∆N L m(t)] is the number of CPs situated ini- tially (at t = 0) in the L (R) aggregate and having, at time t, the x(t) coordinate in the interval {−2R0 + (m − 1)∆x,−2R0 + m∆x}, 1 ≤ m ≤ M , with M a prop- erly chosen, sufficiently large integer, ∆x = 4R0/M , and ∆Nm(t) = ∆N L m(t). Clearly, dmix can take values between 0 (completely unmixed system) and 1 (uniformly mixed system). m(t) + ∆N R The time evolution of dmix(t) is shown in Fig. 5. In the KMC simulation cellular mixing is almost complete (dmix = 1) after the characteristic fusion time τ , i.e., sig- (color online). Comparison of (r/R0)2 vs t/τ for FIG. 3. the fusion of two spherical aggregates obtained from KMC simulations (dashed line), CPD simulations (thin solid line), continuum theory (thick solid line) and experiment (circles) using cardiac cushion tissue (CT) aggregates [22]. 3. Toroidal structure formation Once the KMC and CPD time scales have been cali- brated from the fusion of two spherical CT aggregates, one can employ KMC and CPD simulations to describe and predict the time evolution of more complex CT struc- tures, which are not tractable analytically. To exemplify this point, here we consider the formation of a toroidal structure as a result of the fusion of 10 identical CT spherical aggregates initially arranged in a circular con- figuration as shown in Fig. 4a. The corresponding KMC and CPD simulations were carried out using the same model parameters as in the fusion of two aggregates de- scribed above. In both KMC and CPD simulation the fusion process into a toroidal ring appeared to be com- pleted in ∆t ≈ 2.5τ ≈ 12.5 h, as shown in Fig. 4b. This prediction should be easily testable experimentally. While it seems that both KMC and CPD methods are capable of providing a fairly good description of the shape evolution of a multicellular system during its biomechan- ical relaxation process, the actual cellular dynamics in the two methods is quite different. Indeed, unlike in CPD simulations, in KMC simulations the motion of in- dividual cells is unrealistically fast. This point is man- ifest in Fig. 4. By the time the toroidal ring structure is formed, in the KMC simulation, cells from adjacent aggregates (colored differently) appear to be completely mixed. This is clearly not the case in the CPD simu- lations, where, similarly to existing experimental results [8, 22], there is little mixing between the cells of the fused adjacent aggregates. To further emphasize this point, we have quantified the degree of cellular mixing during the fusion, along the x-axis, of two identical spherical aggregates [labeled as L (left) and R (right)], with initial radius R0 (see Fig. 1), 01234567t/τ0.00.51.01.5(r / R0 )2KMCCPDContinuum TheoryExperimentKMCCPD(a)(b)(c) 8 TABLE I. Values of the model parameters (energies expressed in units of ET ) used in the KMC and CPD simulations shown in Fig. 6. Simulation aa 1.0 1.0 1.0 0.8 0.8 0.8 KMC KMC KMC CPD CPD CPD ab 1.1 1.5 0.3 0.9 1.1 0.2 aa+bb 2 1.4 1.4 1.4 1.0 1.0 1.0 γab Case C1 0.3 -0.1 C2 C3 1.1 C1 0.2 C2 -0.1 0.8 C3 Outcome Fig. 6b left Fig. 6c left Fig. 6d left Fig. 6b right Fig. 6c right Fig. 6d right cal aggregate into two attached homogenous spheroidal caps (each containing either a or b cells) as shown in Fig. 6d. Thus, the degree of cell sorting is enhanced (re- duced) for small (large) values of the adhesion energy ab, compared to the corresponding cohesion energies aa and bb. Note that in terms of the interfacial tension γab (defined below Eq. (2), for "1" = a and "2" = b), case C1 corresponds to γab > 0 and ab > aa, while case C2 corresponds to γab < 0. The inequalities defining case C3 also imply γab > 0. Thus, in a multicellular aggre- gate with two types of cells, in order to have cell sorting (segregation) the corresponding interfacial tension must be positive (i.e., γab > 0). The larger this parameter the more efficient and complete the sorting. The results of our KMC and CPD simulations, pre- sented next, appear to be in good agreement with in vitro experimental findings for these three cases [28]. 1. KMC simulations We have performed three KMC simulations of cell sort- ing starting with a spherical aggregate composed of a random mixture of Na = 3, 589 less cohesive cells of type a and Nb = 2, 362 more cohesive cells of type b (i.e., with aa < bb). Thus, the spherical aggregate had a total of N = 5, 951 cells, and a radius of about 10 cell diam- eters. The values of the model parameters used in the three KMC simulations, corresponding to cases C1, C2 and C3 described above, are listed in Table I. Each KMC simulation was performed up to 105 (non-uniform) time steps, given by Eq. (5), leading to the final configurations shown in Fig. 6b-d. To quantify the degree of cell sorting as a function of time during the KMC simulations, we used a sorting parameter s defined as [43] N(cid:88) i=1 s = 1 N Nti Ni , (17) where N is the total number of cells in the system, and for a given cell i, Ni (Nti) is the number of nearest neighbor cells regardless of their type (of the same type ti as the FIG. 5. Time evolution of the mixing parameter dmix calcu- lated for the fusion of two cellular aggregates from the CPD (solid line) and KMC (dashed line) simulations. nificantly sooner than the completion of the fusion pro- cess (∼ 6τ ). By contrast, in the CPD simulation even at the end of the fusion dmix ∼ 0.2 (cid:28) 1. Based on these results one may conclude that: (i) the cellular dynam- ics that drives aggregate fusion in the KMC simulations is unrealistic (i.e., the system is too liquid-like), and (ii) the CPD model provides a more realistic and atractive approach to describe biomechanical relaxation processes of multicellular systems. B. Cell Sorting in Two Component Aggregates When two populations of cells of different adhesivi- ties are randomly mixed within a multicellular aggregate, they sort such that the more adhesive cells occupy the in- ternal region while being surrounded by the less adhesive cells. Cell sorting has been extensively studied both in vitro [13, 36 -- 38] and in silico [39 -- 41]. According to DAH, the outcome of cell sorting in a two-component multicellular aggregate (composed of two types of cells, labeled 'a' and 'b') depends on the relative magnitude of the corresponding works of cohesion/adhesion needed to separate cells of the same/different types (i.e., aa, bb, and ab), respectively [28]. Here we employ both KMC and CPD simulations (described in Secs. II A and II B) to investigate cell sort- ing in a spherical aggregate of two cell types a and b, with aa < bb. We consider three cases, referred to as C1, C2 and C3, that lead to qualitatively different experimental outcomes [28]. C1: For intermediate adhesion between a and b cells, i.e., aa < ab < (aa+bb)/2, the less cohesive a cells engulf the more cohesive b cells, thus leading to the complete segregation (see Fig. 6b). C2: For strong a -- b adhesion, i.e., (aa + bb)/2 < ab, there is limited sorting and the spherical aggregate remains more or less homogeneously mixed (see Fig. 6c). C3: For weak a -- b adhesion, i.e., ab < aa < bb, the two types of cells completely separate by transforming the initial spheri- 0123456 t/τ0.00.20.40.60.81.0dmixKMCCPD (cid:20) 1 2 smax ≈ 1 N 9 (cid:21) the interface Nti/Ni ≈ 1/2, according to Eq. (17), × ∆N + 1 × (N − ∆N ) = 1 − 1 2 ∆N N . b d/2, and ∆N × (4π/3)(d/2)3 ≈ 4πR2 Furthermore, assuming that cells are distributed uni- formly within the aggregate, one has Nb(d/2)3 ≈ R3 b , i.e., Rb ≈ N 1/3 b ∆R, implying ∆N ≈ 6N 2/3 (∆R/d). Finally, assuming that the thickness of the interfacial layer, separating the seg- regated cell regions, is ∆R = xd, where 2 < x < 3, one obtains b smax ≈ 1 − 3x N 2/3 b N (18) Note that according to Eq. (18), as N → ∞, i.e., for large aggregates, smax approaches unity as N−1/3 (assuming that Na and Nb are of the same order of magnitude). . The time evolution of the sorting parameter, s = s(t), in our KMC simulation corresponding to case C1 is shown in Fig. 7. The insets represent snapshots of the sorting process taken at times indicated by the arrows. The sharp increase of s(t) at the beginning of the sim- ulation followed by a slow asymptotic approach to smax indicates that there are at least two sorting time scales. Indeed, the entire time evolution of the sorting parameter can be well fitted with the double exponential −t/τ2, s(t) = smax − s1e −t/τ1 − s2e (19) FIG. 6. 3D snapshots from KMC (left) and CPD (right) sim- ulations of the cell sorting in an initially spherical aggregate composed of two randomly mixed cell types (black and light grey). The snapshots represent the (a) initial, and (b-d) final configurations of the simulated system. The latter correspond to (b) intermediate (case C1), (c) strong (case C2), and (d) weak (case C3) cell adhesion energy, as explained in the text. For better visualization of cell mixing/sorting in (a)-(c) only half of the spherical aggregate is shown. (Images rendered with VMD [42]). cell i). The sum in Eq. (17) runs over all cells in the sys- tem. Clearly, 0 < s < 1, and the larger s the more com- plete the sorting. Note that even for completely sorted multicellular systems, built from two (or more) different cell types, the presence of the interface(s) between the segregated regions renders the maximum possible value, smax, of the sorting parameter smax < 1. For example, in the above case C1, when at the end of sorting Na cells of type a completely engulf Nb cells of type b, one can es- timate smax as follows. For simplicity, assume that both cell types have spherical shape with the same diameter d. Let ∆N be the number of cells (of either type a or b) situated at the spherical interface, of mean radius Rb and width ∆R, between the two segregated regions (see Fig. 6b), and N = Na + Nb. Since for a cell i situated at where smax = 0.76 is in very good agreement with the theoretically estimated value 0.78 obtained from Eq. (18) for x = 2.5. The other fitting parameters in Eq. (19) are: τ1 = 1.4t0, s1 = 0.27, τ2 = 58.5t0 and s2 = 0.11. The shorter time scale τ1 corresponds to the local rearrange- ment (sorting) of cells leading to small clusters of same types of cells, while the longer time scale τ2 describes the much slower engulfment process of the b cells by the a cells, a process that requires large displacements by a finite number of cells. Although the results of our KMC simulations appear to be in good qualitative agreement with experiments on cell sorting [12, 13, 28], a quantitative comparison, e.g., in terms of the time evolution of the sorting parameter, is not feasible because s(t) cannot be measured experimen- tally. Thus, there is no simple way to reliably calibrate the time unit t0 (which is related to the model parame- ter w0) used in the plot of s vs t/t0 in Fig. 7. However, s(t) can also be determined from CPD simulations, thus allowing for a quantitative comparisson between the two computer simulation methods. 2. CPD simulations We have also used CPD simulations to investigate cell sorting corresponding to the three cases C1, C2 and C3 KMCCPD(a)(b)(c) (d) 10 good agreement with the theoretical prediction Eq. (18), i.e., 0.67 for x = 2.2 (or 0.63 for x = 2.5). The other fitting parameters in Eq. (19) are: τ1 = 0.68 t0, τ2 = 103 t0, s1 = 0.25 and s2 = 0.1. Note that while s1 and s2 have essentially the same values for both KMC and CPD simulations, the time constants τ1 and τ2 are quite different, as the corresponding time units t0 are different in the two simulations. Moreover, the fact that, for similar model parameters, τ2/τ1 = 41.8 in KMC is about twice as large as τ2/τ1 = 21.7 in the corresponding CPD simulation indicates that the self diffusive motion of cells in KMC occurs much faster than in CPD. In other words, the multicellular system is more liquid-like in KMC than in CPD simulations. IV. CONCLUSIONS We have presented two (KMC and CPD) simulation methods and an analytic, continuum theoretical ap- proach to address structure formation by cell sorting and the fusion of contiguous multicellular spheroids. The the- oretical method was used to interpret the experimental results, and to test the viability and calibrate the model parameters of the KMC and CPD simulations. Our study was motivated by the need to quantify biomechanical properties of engineered tissue constructs, composed of compact tissues made of adhesive and motile cells and to predict their time evolution. The growing interest for understanding shape changes in such tissue constructs stems from their applications in tissue engineering in gen- eral and in the emergent field of 3D bioprinting in par- ticular [8]. The KMC method is based on a lattice representa- tion of the 3D tissue construct and dynamics is described in terms of rates associated with possible movements of cells. Similarly to previously employed MMC studies, the mixing pattern observed in KMC simulations disagrees with experiments. In both methods an elementary move consists in cells swapping positions with neighbors, which overestimates cell motility. Nevertheless, the time scale calibration in KMC makes the simulated time course re- alistic as far as the shape evolution of multicellular tissue constructs is concerned. The CPD method is based on modeling individual cells in a tissue construct as interacting CPs. The dynamics of the multicellular system are determined by integrat- ing the equations of motion for each CP. The CPD force field parameters are determined such that the time evolu- tion of the shape of the fusing spherical aggregates in the CPD simulation matches as closely as possible the exper- imental one (i.e., two touching spherical caps). Once the CPD model is calibrated, this can be used to simulate the shape evolution of arbitrary 3D multicellular constructs. It should be emphasized that in CPD (i.e., computer) units the calibrated CPD parameters (and therefore the outcome of a CPD simulation) are independent of the used cell type. However, the CPD units (9) have spe- FIG. 7. Time dependence of the cell sorting parameter, s = s(t), corresponding to case C1 described in the text, for both KMC (top) and CPD (bottom) simulations. The insets represent snapshots of half of the spherical aggregate taken at times indicated by arrows. a ab b b = intra a = intra = 0.8 and bb ≡ inter described above. The initially spherical aggregate con- tained a random mixture of equal number Na = Nb = 1, 000 of cells of type a and b. While the CPD parameters aa ≡ inter = 1.2 were kept the same in all three simulations, the parame- ter ab ≡ inter had different values (similar to the ones used in the KMC simulations) for the three cases C1, C2 and C3 as listed in Table I. The cell sorting patterns obtained at the end of the corresponding CPD simula- tions are shown in Fig. 6. As expected, these patterns are similar to the ones obtained in the KMC simulations. In order to quantify the degree of cell sorting in the CPD simulations by employing the cell sorting parameter s, defined through Eq. (17), we determined the position of a cell by the center of mass of the constituent CPs, and considered two cells to be neighbors if they were separated by a distance less than 3.25 σ. For the CPD simulation corresponding to case C1, s(t) is shown Fig. 7. Similarly to the KMC result, s(t) can be fitted well with the double exponential (19). Again, smax = 0.68 is in 050100150200t/t00.30.40.50.60.70.80.91.0s0100020003000t/t00.30.40.50.60.70.80.91.0sKMCCPD 11 cific values for different cell types. Thus, the CPD sim- ulations reported here can be applied as is to different cell types; the corresponding CPD time unit t0 should be determined in each case by equating τsim ≈ 540 t0 (see Sec. III A 2) with the experimental fusion time τexp. The reported CPD simulations provided a good de- scription for both fusion and cell sorting of multicellular spheroids. We found that CPD provides a more realistic description of complex multicellular structure formation than KMC. Indeed, the behavior of the studied multi- cellular systems in CPD simulations resembles to that of complex visco-elastic materials while in KMC simula- tions to that of viscous liquids. It is to be expected that by including more realistic features into the interaction of the CPs the accuracy of the CPD method can be further improved. ACKNOWLEDGMENTS This work was supported by grants from the National Science Foundation (PHY-0957914 and FIBR-0526854). The work by A.N. was supported in part by the Roma- nian National Authority for Scientific Research (CNCSIS Contract PCCE-ID 76). Computational resources were generously provided by the University of Missouri Bioin- formatics Consortium. [1] D. Ingber and M. Levin, Development 134, 2541 (2007). [2] G. Forgacs and S. Newman, Biological physics of the de- veloping embryo (Cambridge University Press, 2005). [3] R. Langer and J. P. Vacanti, Science 260, 920 (1993). [4] L. G. Griffith and G. Naughton, Science 295, 1009 (2002). Association of Anatomists 237, 2438 (2008). [23] J. Frenkel, J. Physics 9, 385 (1945). [24] O. Pokluda, C. T. Bellehumeur, and J. Vlachopoulos, Aiche Journal 43, 3253 (1997). [25] C. T. Bellehumeur, M. Kontopoulou, and J. Vlachopou- los, Rheologica Acta 37, 270 (1998). [5] V. Mironov, T. Boland, T. Trusk, G. Forgacs, and R. R. [26] A. B. Bortz, M. H. Kalos, and J. L. Lebowitz, J Comput Markwald, Trends Biotechnol 21, 157 (2003). Phys 17, 10 (1975). [6] K. Jakab, A. Neagu, V. Mironov, R. R. Markwald, and G. Forgacs, Proc Natl Acad Sci U S A 101, 2864 (2004). and X. Cui, Biotechnol J 1, 910 [7] T. Boland, T. Xu, [27] J. Israelachvili, Intermolecular and Surface Forces (Aca- demic Press, 1997). [28] R. A. Foty and M. S. Steinberg, Int J Dev Biol 48, 397 (2006). (2004). [8] K. Jakab, C. Norotte, B. Damon, F. Marga, A. Neagu, C. L. Besch-Williford, A. Kachurin, K. H. Church, H. Park, V. Mironov, R. R. Markwald, G. Vunjak- Novakovic, and G. Forgacs, Tissue Engineering: Part A 14, 413 (2008). [9] V. Mironov, R. P. Visconti, V. Kasyanov, G. Forgacs, C. J. Drake, and R. R. Markwald, Biomaterials 30, 2164 (2009). [10] C. Norotte, F. S. Marga, L. E. Niklason, and G. Forgacs, Biomaterials 30, 5910 (2009). [11] K. Jakab, C. Norotte, F. Marga, K. Murphy, G. Vunjak- Novakovic, and G. Forgacs, Biofabrication 2, 1 (2010). [12] M. S. Steinberg, Science 141, 401 (1963). [13] M. S. Steinberg, J Exp Zool 173, 395 (1970). [14] A. Neagu, K. Jakab, R. Jamison, and G. Forgacs, Phys- [29] D. A. Beysens, G. Forgacs, and J. A. Glazier, Proceed- ings of the National Academy of Sciences of the United States of America 97, 9467 (2000). [30] M. S. Steinberg and T. J. Poole, "Liquid behavior of em- bryonic tissues," in Cell Behavior, edited by R. Bellairs, A. S. G. Curtis, and G. Dunn (Cambridge University Press, 1982) pp. 583 -- 697. [31] S. P. Palecek, J. C. Loftus, M. H. Ginsberg, D. A. Lauf- fenburger, and A. F. Horwitz, Nature 385, 537 (1997). [32] M. H. Zaman, R. D. Kamm, P. Matsudaira, and D. A. Lauffenburger, Biophysical Journal 89, 1389 (2005). [33] S. Plimpton, Journal of Computational Physics 117, 1 (1995). [34] R. Gordon, N. S. Goel, M. S. Steinberg, and L. L. Wise- man, J Theor Biol 37, 43 (1972). ical Review Letters 95, 178104 (2005). [15] R. A. Foty, G. Forgacs, C. M. Pfleger, and M. S. Stein- [35] J. D. Eshelby, Trans. AIMME 185, 806 (1949). [36] M. S. Steinberg, Proc Natl Acad Sci U S A 48, 1577 berg, Physical Review Letters 72, 2298 (1994). (1962). [16] R. A. Foty, C. M. Pfleger, G. Forgacs, and M. S. Stein- berg, Development 122, 1611 (1996). [37] M. S. Steinberg, Exp Cell Res 30, 257 (1963). [38] J. M. Perez-Pomares and R. A. Foty, Bioessays 28, 809 [17] E. Flenner, F. Marga, A. Neagu, I. Kosztin, and G. For- (2006). gacs, Curr. Top. Dev. Biol. 81, 461 (2008). [39] F. Graner and J. A. Glazier, Physical Review Letters 69, [18] T. J. Newman, Mathematical Biosciences and Engineer- 2013 (1992). ing 2, 611 (2005). [40] J. A. Glazier and F. Graner, Physical Review E 47, 2128 [19] S. A. Sandersius and T. J. Newman, Phys Biol 5, 15002 (1993). (2008). [41] J. C. Mombach, J. A. Glazier, R. C. Raphael, and M. Za- [20] S. A. Sandersius, C. J. Weijer, and T. J. Newman, Phys- jac, Physical Review Letters 75, 2244 (1995). ical biology 8, 045007 (2011). [21] S. A. Sandersius, M. Chuai, C. J. Weijer, Newman, Physical biology 8, 045008 (2011). [42] W. Humphrey, A. Dalke, and K. Schulten, Journal of and T. J. Molecular Graphics 14, 33 (1996). [43] E. Palsson, Future Generation Computer Systems 17, [22] K. Jakab, B. Damon, F. Marga, O. Doaga, V. Mironov, I. Kosztin, R. Markwald, and G. Forgacs, Developmen- tal dynamics : an official publication of the American 835 (2001).
1504.04430
1
1504
2015-04-17T02:58:43
Critical Waves and the Length Problem of Biology
[ "physics.bio-ph", "q-bio.SC" ]
It is pointed out that the mystery of how biological systems measure their lengths vanishes away if one premises that they have discovered a way to generate linear waves analogous to compressional sound. These can be used to detect length at either large or small scales using echo timing and fringe counting. It is shown that suitable linear chemical potential waves can, in fact, be manufactured by tuning to criticality conventional reaction-diffusion with a small number substances. Min oscillations in E. coli are cited as precedent resonant length measurement using chemical potential waves analogous to laser detection. Mitotic structures in eucaryotes are identified as candidates for such an effect at higher frequency. The engineering principle is shown to be very general and functionally the same as that used by hearing organs. PNAS Significance Statement: This paper invokes physical principles to address the question of how living things might use reaction-diffusion to measure out and regulate the many thousands of lengths required to make their body parts and internal organs. It argues that two ideas have been missing. One is that oscillation is necessary to achieve the necessary design stability and plasticity. The other is that the system must be tuned to criticality to stabilize the propagation velocity, thus enabling clocks to function as meter sticks. The broader significance is twofold: First, a fundamental piece of the machinery of life is probably invisible to present-day biochemical methods because they are too slow. Second, the simplicity of growth and form identified a century ago by D'Arcy Thompson is probably a symptom of biological engineering strategies, not primitive law.
physics.bio-ph
physics
Critical Waves and the Length Problem of Biology Department of Physics, Stanford University, Stanford, CA 94305∗ Robert B. Laughlin (Dated: March 1, 2015) It is pointed out that the mystery of how biological systems measure their lengths vanishes away if one premises that they have discovered a way to generate linear waves analogous to compressional sound. These can be used to detect length at either large or small scales using echo timing and fringe counting. It is shown that suitable linear chemical potential waves can, in fact, be manufac- tured by tuning to criticality conventional reaction-diffusion with a small number substances. Min oscillations in E. coli are cited as precedent resonant length measurement using chemical potential waves analogous to laser detection. Mitotic structures in eucaryotes are identified as candidates for such an effect at higher frequency. The engineering principle is shown to be very general and functionally the same as that used by hearing organs. It is not known how living things measure their lengths. This is true notwithstanding the immense progress made over the past 30 years in understanding morphogen gra- dients in embryogenesis.1 -- 6. The problem is captured nicely by the confusion over regulation of the bicoid pro- file in Drosophila7 -- 11, but it is also reflected in the noto- rious instability, hysteresis, and lack of scalability of tra- ditional static reaction-diffusion12,13. No one knows why cells are the size they are14, why plants and animals are the size they are15, how organs grow maintaining their proportions16, and how some animal bodies regenerate lost limbs17. On the matter of length determination, per se, very little progress has been made beyond Thomp- son's 1917 treatise on biological form18. Length has a special place in biology by virtue of be- ing a primitive quantity with units. It is not possible for living things to size themselves properly without hav- ing developed the skill of measuring these quantities as numbers and relating these numbers to each other math- ematically. They require meter sticks to do this. They must fabricate these meter sticks using diffusion and mo- tors, since they are the only biochemical elements that involve length. The relationships of these meter sticks to each other and to the lengths they measure must be pre- cise and described by equations. This is because precise mathematical relationships among lengths are what size and shape are. In this paper I point out that the difficulty of ac- counting for length relationships of parts of organisms with equations disappears instantly if the organism is premised to have discovered a way to emulate elemen- tary physical law. In particular, one simple invention is sufficient to facilitate the measurement and construc- tion of body plans of any shape and size one might wish in a way that is both plastic and scalable: the conver- sion of diffusive motion into linear waves using engines. The concept is general because all motion in the pres- ence of disorder, including motion of cytoskeletal compo- nents, becomes diffusive at long time and length scales by virtue of evolving into a random walk. But if engines can transform this random walking into propagation with stable direction and speed, then signals can be beamed, like a flashlight, reflected from boundaries, and trapped. FIG. 1: Chemical amplifier described in electrical terms. The circuit components Rg, Lg and Cg all have negative values. On the right is shown the impulse response described by Eq. (1) for the special case of RgCg = Lg/Rg = τ . The voltage step height V0 is set to 1. The current response, plotted as a multiple of V0/Rg, is negative and has the same shape as a neuron action potential. Once this happens, the organism can measure lengths the same way human engineers do, by echo timing or by fringe counting and resonance. Body designs based on this strategy are inherently plastic because fixing the speed enables lengths to be laid out or detected by means of clock tick intervals, which are easy to change. Although it is not widely known, biological systems can easily manufacture such waves using elementary reaction- diffusion chemistry similar to that at work in neuron ac- tion potentials.19 The key is tuning the chemical reac- tions to the edge of an instability, an effect known in the cochlear amplifier literature as criticality.20 -- 22 As I shall show, this trick is so easy to implement technically that it is hard to imagine how Nature would not have exploited its tremendous engineering advantages in the struggle for survival. This obligates us to take seriously even very small hints that Nature did, in fact, discover how to do it long ago. The simple explanation for why we have found only sparse empirical evidence for such waves so far is that chemical potential waves are difficult to detect with existing laboratory techniques. The larger idea implicit in this proposed resolution of the length problem is that biological systems cannot conduct en- gineering without rules any more than we humans can, 5 1 0 2 r p A 7 1 ] h p - o i b . s c i s y h p [ 1 v 0 3 4 4 0 . 4 0 5 1 : v i X r a so they invented some in the ancient past, and the ones that worked best turned out to be same ones we humans discovered later using reason. I. CHEMICAL POTENTIAL WAVES The simplest chemical length mensuration apparatus involves chemical potential waves solely. Apparati with other components, such as mechanical motors, are al- lowed also, but all of them necessarily have a chemical potential component by virtue of how they work. Despite being difficult to detect, chemical potential waves are known to be pervasive in biology. The most familiar case is the neuron action potential, the electri- cal aspects of which make it easy to detect by primi- tive means, even though it is fundamentally a reaction- diffusion wave23. But there are also non-electrical va- rieties: cAMP waves in slime molds, which direct the colony's organization into fruiting bodies24; calcium waves, directly implicated in oogengenesis25,26, develop- mental patterning27, brain function28 -- 30, and cell signal- ing in animals31 and plants32; and MinDE waves in E. coli, perhaps the most important of all because they are involved in a bacterial length decision33. All of these non-electrical versions require highly advanced technolo- gies to see, and also required a bit of luck to find, so there is good reason to suspect that more exist and simply have not been detected yet. The simplest way chemical reactions can manufacture waves is through stable two-terminal amplification, the fundamental basis of laser operation34. The observation that amplification is involved is important, for while all amplifiers exploit nonlinearities to work, there is nothing inherently nonlinear about what they do. All amplifiers become nonlinear when they are pushed to deliver large powers. It is thus not surprising that the wave signals easiest to observe in biology are often nonlinear. But amplification in the linear regime is known to occur as well, notably in hearing organs35 -- 40. 2 FIG. 2: Left: Illustration of a diffusive transmission line with a series of amplifiers like those in Fig. 1 placed across it. R and C represent the resistance and capacitance per unit length before the amplifier is added. The repeat distance is b. Right: Plot of the solution of Eq. (4) for the special case of C + Cg = 0 and RgCg = Lg/Rg = τ . Both ω and qv are expressed as multiples of 1/τ . The complex numbers ω and its negative complex conjugate −ω∗ are two of the three roots, the third being pure imaginary and off the top of the graph. The dashed line is the solution when the amplifiers are turned off (Cg → 0). a molecular stopper43. The Lg < 0 mainly causes a fi- nite turn-on time Lg/Rg, but it also adjusts the circuit's after-bounce, so it corresponds to the K+ channel of a neuron. Thus Fig. 1 is simply a linearized version of the Hodgkin-Huxley equations. It is important that both Cg < 0 and Lg < 0 are dynamical creations of the amplifier itself, not additional postulates. It is physically impossible to make a stable amplifier without also creating, as a side effect, negative reaction. This effect is seen in lasers as a reversal of the dielectric function whenever the laser gain medium becomes amplifying44; but the deeper reason has nothing to do with quantum mechanics or population inversions. It is causality45. The current induced by a stimulating voltage can appear only after the stimulus is applied, never before. The current induced by the step voltage shown in Fig. 1 is (cid:90) ∞ −∞ II. AMPLIFIERS: STABILITY AND CAUSALITY I(t) = V0 2π Cge−iωt 1 − iωRgCg − ω2LgCg dω (1) Two-terminal amplifiers are easiest to explain by elec- trical analogy. Consider the circuit shown in Fig. 1. It is a conventional linear resonator, such as one might find in any radio, except that the components all have neg- ative values. Its active component, the negative resis- tor, causes electric current to flow in a direction opposite to the way it normally would when voltage is applied. Such reversed flow is implicit in all Na+-K+ neural mod- els, including the original one of Hodgkin and Huxley41, but it is quite explicit in those based on tunnel diodes42. The all-important Cg < 0 causes induced current to stop flowing after a time RgCg. Na+ channels achieve this cessation by plugging themselves after a time delay with It is properly causal provided that poles of the response kernel lie in the lower half of the complex plane. This requires both Cg and Lg to be negative if Rg is. The physical principles operating in Fig. 1 apply to all amplifiers, not just electrical ones. The denominator in Eq. (1) may be seen to be a Taylor expansion in ω truncated to second order. Such a truncation is always valid at long times, and it is equivalent to stating that the system has only two poles in the complex plane and thus has only two important mechanical degrees of freedom. Those things are therefore not model assumptions at all but generic features of amplifier response at long times. In the case of Fig. 1, the degrees of freedom are charge and current, but in general they could be anything. III. TURING CRITICAL WAVEFUNCTION Linear waves are produced when one places a series of such amplifiers across a diffusive transmission line, as shown in Fig. 2. This is aptly analogous to placing Na+- K+ amplifiers across the membrane of an axon. Substi- tuting Vj = V0 exp[i(qbj − ωt)], where b is the repeat length, for the voltage on the jth site, we obtain the dis- persion relation 3 [1 − cos(qb)] − iωC − iωCg 2 R When the amplifiers are turned off (Cg → 0), this be- comes the diffusion equation 1 − iωRgCg − ω2LgCg = 0 (2) Dq2 − iω = 0 (D = b2 RC ) (3) in the limit of small q. But when the amplifiers are turned on, and also adjusted so that C+Cg = 0, Eq. (2) becomes (vq)2 − ω2(1 − iωτ ) 1 − iωτ − ω2τ 2 = 0 This is functionally equivalent to the wave equation (v =(cid:112)D/τ ) (v =(cid:112)D/τ ) (4) (5) ∂2ψ ∂x2 = 1 v2 ∂2ψ ∂t2 where ψ is any one of the dynamical variables, in the regime qv < 0.5/τ . The expression for the velocity v is the Luther equation46. The full solution Eq. (4) is plotted in Fig. 2. This wave equation is equivalent to the Turing reaction-diffusion equations FIG. 3: Left: Solutions of Eq. (7) with Neumann bound- ary conditions for a pill-shaped cavity, as appropriate for a bacterium. Hemispherical end caps of radius a are attached to a cylinder of length 2.2 a. The top shows a contour plot of the lowest eigenfunction, which corresponds to an axially symmetric pole-to-pole sloshing. It occurs at ωn = 0.88v/a. Below that are contour plots of first longitudinally symmetric mode, which corresponds to a steady migration around the perimeter along a path perpendicular to the axis. It occurs at ωn = 1.91v/a. This mode is degenerate with a mirror- reflected one that rotates in the opposite direction. Right: Solution of Eq. (7) for a spherical cavity with mixed bound- ary conditions (n·∇φn = −0.51φn). The top shows a contour plot of the lowest eigenfunction. The dotted lines, reproduced as solid lines below, are trajectories perpendicular to the con- tour lines. having the right mix of Na+ and K+ channels. The other is cancellation of the capacitances, a result achieved in practice by slowly increasing the number of amplifiers in the membrane until the system begins to oscillate a little. Such oscillations are routinely observed emanating from the ear47. A similar phenomenon has been reported at the surface membranes of yeast48. ∂X ∂t ∂Y ∂t ∂Z ∂t = = = 1 τ 1 τ 1 τ Z + D ∂2X ∂x2 Z (X − Y − Z) IV. LENGTH MEASUREMENT USING CAVITY RESONANCE (6) It could thus easily be achieved with chemical reactions among three substances. It is also easily generalized to three dimensions. In fact, the principle behind Eq. (4) is so general that it applies to any conservative diffusive phenomenon, chemical, electrical or mechanical, at any length or time scale, regardless of details. For this reason, it is a competitive candidate for how living things might measure their lengths generally. Stabilization of the wave velocity in these systems is achieved through two fine-tunings. One is equality of the capacitive and inductive times. This is a matter of amplifier design and is achieved in the case of neurons by The simplest strategy for measuring length with man- ufactured waves is detecting mode resonances in a cav- ity. This is illustrated in the case of 3-dimensional waves in Fig. 3. The allowed oscillations of a cavity are found by substituting a harmonic solution ψ(r, t) = φn(r) exp(−iωnt) into Eq. (5) and solving the Helmholtz equation ∇2φn + (ωn/v)2 φn = 0 (7) with Neumann boundary conditions (n · ∇φn = 0), as appropriate for a substance that is conserved and cannot flow in or out through the walls. Solutions exist only for certain discrete eigenfrequencies ωn. The values of these and the spatial behavior of their corresponding eigen- functions φn sense the size and shape of the cavity. The eigenfunction φn corresponding to the lowest of eigen- frequency shown in in Fig. 3 describes to the observed MinDE wave motion in E. coli, although details differ. 4 (cid:26) (cid:20) 1 + f0 1 Rg → 1 Rg gC(cid:48) (cid:21)(cid:27) (8) −iωτ0 1 − iωτ0 − (ω/ω0)2 gC(cid:48) To actually excite eigenmodes of a cavity it is necessary provide the amplifying medium with a gain peak. This is achieved most simply in the case of the example of Fig. 2 by attaching a second resonant circuit, as shown in Fig. 4. This effectively makes the amplifying resistor slightly frequency dependent, per g, ω0 = (L(cid:48) g)−1/2 and f0 = Rg/R(cid:48) where τ0 = R(cid:48) g. Modifying Rg in this way is physically equivalent to hum- ming a tone in a closed room: The tone is ω0, the loudness is f0, and the time between successive breaths is τ0. The corresponding reaction-diffusion equations are dX dt dY dt dZ dt dY (cid:48) dt dZ(cid:48) dt = D Z 1 τ ∂2X ∂x2 + 1 Z τ (X − Y − Z + f0Z(cid:48)) 1 τ Z(cid:48) 1 τ0 (cid:2)Z − Y (cid:48) − (1 + f0)Z(cid:48)(cid:3) = = = = ω2 0τ0 (9) When modified in this way the system becomes a text- book laser oscillator. As shown in Fig. 4, cavity modes with frequencies ωn in the region of net gain grow ex- ponentially and saturate the amplifier, meaning they eat up all the power available. In biological terms we would say that a nonlinearity chokes off the exponential growth and causes it to plateau. Laser saturation reduces the gain according to the approximate formula f0 −→ f = f0 1 + P/P0 (10) where P is the power delivered and P0 is a parameter characteristic of the medium. P continues to increase until Im(ω) becomes zero, as shown in Fig. 4, for the mode with the highest native gain. This implies that all the other modes in the saturated state have negative gain and die away. This winner-take-all competition for the available energy causes the system to oscillate in one mode rather than many. The frequency and magnitude of the saturated oscilla- tion both measure the cavity length. The coarse-grained FIG. 4: Left: Illustration of the modification of Rg required to introduce a gain peak into the transmission line of Fig. 2, as described by Eq. (8). Lower Right: Correction to dispersion relation of Fig. 2 resulting from the values ω0 = 0.2/τ , f0 = 0.1, and τ0 = 10τ . Only the imaginary part of ω is shown because the correction to the real part is negligible. Hatching indicates the region of oscillation. Here saturation, described by Eq. (10), pushes Im(ω) to zero. The dashed line shows the g → 0 behavior. Both ω and qv are expressed as multiples C(cid:48) 1/τ . Upper Right: Power produced at saturation, per Eq. (10). measurement is the discrete frequency jumping that oc- curs as one mode after another becomes dominant as the cavity is lengthened. The power P of the victorious mode is also modified by the length adjustment through the medium's gain profile, as shown in Fig. 4. The saturated power thus provides a fine-tuning measurement of length. V. PRECEDENT IN BACTERIA Gain, oscillation and saturation have all been observed experimentally for MinDE oscillations in E. coli.49. Flu- orescence tagging experiments have revealed that MinD molecules flock together from one end of the bacterium to the other with a round-trip travel time of about 40 seconds. Disabling expression of FtsZ, a protein required for septation and division, causes the bacterium to grow very long and exhibit a preferred MinD wavelength of ap- proximately 10 µm, or about twice the length at which it normally divides49. Both standing waves and travel- ing waves are observed in these long mutants, depending on circumstances.49,50 The system can also flip unsta- bly between the two when the boundary conditions are changed. When the bacterium divides, the oscillation bi- furcates unstably and then settles down with a higher frequency, just as a laser would51. The increase is mea- sured to be a factor of 1.5, whereas a factor of 2 would be expected of perfectly linear waves. There is no direct evidence that the bacteria oscillate for the purpose of measuring their lengths absolutely, nor is there any direct evidence that oscillations are univer- sally present in all bacteria. One knows for certain only that E. coli use Min oscillations are part of the machin- ery for determining their midpoint for division, and that a MinD homolog in B. subtilis has been observed to form static patterns that do not oscillate.52 However, circumstantial evidence is abundant. The existence of Min oscillations clearly demonstrates that chemical reactions actually present in a bacterium have the ability to manufacture chemical waves and trap them. Nature presumably created this machinery for some pur- pose. Min oscillations are known to involve only a small number of substances. The exact number is controver- sial, but the reactions are typically modeled with four or five, the same as in Eqs. (9).53 -- 57. Existing experi- ments are not sufficiently detailed to distinguish among these models, but elementary reaction-diffusion is cen- tral to all of them, and all become similar to Eqs. (9) when linearized. The fundamental simplicity of the reac- tions was indicated early on by identification of MinCDE operon damage as the cause of the minicell mutation in E. coli58,59, but it is now corroborated by experiments in vitro showing both oscillations and spatial waves oc- curring in system containing only MinD, MinE, a lipid membrane, and ATP60 -- 62. The frequencies and lengths observed in these experiments do not agree with those observed in real bacteria, but this is not surprising given how finely tuned a reaction-diffusion system must be to measure lengths accurately. They are like a watch with a corrupted regulator: It still ticks, but it does not keep time. It is not important that the Min amplifier machinery resides in or near the cell membrane63,64. For length measurement purposes, this machinery is adiabatically equivalent to Eq. (5), meaning that it can be slowly deformed into scalar waves trapped in the bacterial body without changing its functionality. The corresponding wave speed v is about 0.15 µm/sec, the same as slow calcium waves. The wave principle also has potential bearing on the overall shape of bacteria, most of which are cylinders of fixed width. The reasons for preferring this shape are not presently known.65 To measure the width of the body with a wave, one must excite the first azimuthal eigen- mode of Eq. (5), also shown in Fig. 3. This corresponds to a motion around the perimeter perpendicular to the body axis. This mode is necessarily doubly degenerate so long as the body is exactly cylindrical, so the correspond- ing form would not automatically be a cylinder unless the symmetry is broken, meaning that either right-handed or left-handed motion is preferred. This is a different issue from the handed spiral structures reported in bacterial walls66 because it requires also breaking of time-reversal symmetry, as occurs in a magnet. Such symmetry break- ing is known to occur in E. coli, where it manifests itself through preferred swimming handedness on glass slides, an effect attributable to a preferred rotation direction of the flagellum.67,68 The shape-regulating protein MreB has recently been observed to execute motion circumferential and perpen- dicular to the body axis in B. subtilis69 -- 71. The ex- periments employ difficult sub-wavelength optical mi- 5 FIG. 5: Illustration of standing wave syncytium. Top: Plot of the time average of ψ(x, t)2 = cos(ωx/v) cos(ωt) +  sin(2ωx/v) cos(2ωt)2 at fixed frequency ω and wave speed v for two different values of . This shows how a standing wave grating can be made to split centrosome locations by increas- ing the amplitude of its first harmonic. Middle: The same quantity for the two-dimensional case expressed as a contour plot. The wavefunction is a "twinkling eyes" superposition of three plane waves at frequency ω and three more at 2ω.74 Bot- tom: Contour in the hexagonal unit cell showing phase-locked signal contours of ψ and spindles constructed perpendicular to them, as in Fig. 3. croscopy, and some details remain controversial. The re- ported azimuthal velocities range from 7 nm/sec to 50 nm/sec, and one group reports a handedness bias while another rules it out. However, there is general agreement that MreB aggregates into small patches and that these translocate along the cell wall in a direction accurately perpendicular to the body axis. In the context of mensuration it is not important whether the patch motion involves cell wall synthesis, as the experiments seem to suggest it does. The principles by which diffusion is converted to wave motion are so general that they apply equally well to polymerization. VI. EUCARYOTES: SPINDLES AND SYNCYTIA Anything one says about eucaryotic size and shape control is necessarily speculative because so little defini- tive is known about it. However, the physical principles of resonant trapping are so simple and general that one might reasonably guess that they apply also to eucaryotic cells. Fig. 3 also shows the solution of Eq. (7) for a spheri- cal cavity, as might be appropriate for a eucaryotic cell. Everything is the same as for the pill-shaped cavity ex- cept for the boundary conditions, which we force to be mixed, thus pushing the antinode from the cell surface into the interior. Mixed boundary conditions are appro- priate for amplification machinery that resides partly in bulk interior and partly on or near the membrane. Fig. 3 also shows trajectories generated by the rule of every- where going downhill in the wavefunction gradient. The similarity to the mitotic spindle is unmistakable. Thus were an oscillating chemical potential field providing the navigation instructions for microtuble assembly, it would account quantitatively for (1) the location of centrosomes (the antinodes), (2) the existence and location of the metaphase plate, (3) the choice of a particular orien- tation for this plate, (4) the initiation and termination microtubules at the centrosomes, (5) their intersection at right angles with the metaphase plate, (6) their out- ward bulging at the plate, (7) the oblique angle formed between backward-going microtubules and the cell mem- brane, and (8) the observed scaling of the spindle assem- bly with cell size72,73. Waves also have the potential to account for the or- ganization of structures without cell membranes. Fig. 5 shows a simple model of a syncytium made with stand- ing waves. This specific construction uses three waves oriented at a physical angle 2π/3 with respect to each other and also oscillating 2π/3 out of phase with each other in time so as to create a "twinkling eyes" dynami- cal pattern.74 The recipe for locating the spindles is the same as in Fig. 3. The equations used to generate Fig. 5 are much too primitive to describe an actual syncytium, among other reasons because the nuclei in real syncytia are not (cannot be) hexagonally arranged and because they have cytoskeletal structures where the cell mem- branes would normally have been. Nonetheless Fig. 5 shows how standing waves can create organizational pat- terns beyond the immediate neighborhood of a specific nucleus, and thus how they might organize larger multi- cellular eucaryotic structures. There are some additional potential benefits, such as providing a natural signal to synchronize mitotic division and automatically scaling structures in a syncytial embryo to egg size. VII. CONCLUSION: STATIC VERSUS DYNAMIC REACTION-DIFFUSION Dynamic reaction-diffusion is not a novel length men- suration method so much as an engineering advance over an older, more primitive one. When implemented at the molecular level, it uses exactly the same chemistry that static reaction-diffusion does but simply manages time differently. An apt analogy would be the time manage- ment that distinguishes the Internet from the telegraph. Both use electricity to work, but the latter uses it more cleverly and is thus vastly more powerful. Dynamic men- 6 suration is thus fully compatible with experimental evi- dence that small organisms use the static version often if the latter is imagined to be vestigial.75 -- 77 The crucial engineering advantage of dynamic mensu- ration over the static variety is plasticity. To measure out a length with an elementary diffusive morphogen one must balance a uniform destruction rate against a dif- fusion constant, a strategy that works perfectly well so long as the design is fixed. But if one wishes to change the design, one must modify the diffusion constant, the destruction rate, or both, making sure that the latter remains uniform. If, on the other hand, one tunes the chemical reactions to manufacture waves with a fixed speed, lengths can be easily adjusted up or down simply by changing frequencies of stimulating oscillators. These need not be located in any particular place, for stand- ing waves are rigid and thus insensitive to the location of their stimulus, an effect familiar from the operation of musical instruments. Thus the difficult hardware design need only be done once. The hardware can then be used again and again to measure out lengths of any size one likes, even with the latter changing on the fly in response to external events not encrypted in the genes. The other important advantage of dynamic mensu- ration is generalizability. Once the concept of turning diffusion into waves using amplifiers is discovered, it is very easy to imagine going by small steps to the inven- tion of a sophisticated organ like the cochlea, which uses the same engineering principles but exploits mechanical diffusion, not chemical diffusion. It is similarly easy to imagine going by small steps to the invention of neurons, which involve the same circuitry but with the amplifier gain turned up to make the propagating pulse nonlinear, and in which the underlying diffusive motion is electrical, not chemical. Diffusion is a very general physical phe- nomenon that results when when motion becomes disor- ganized. The trick of reversing the descent into diffusive chaos using engines thus has applicability far beyond ba- sic chemistry. It is not a great concern that direct biochemical evi- dence for dynamic length measurement is thin. Bacteria, in particular, are very ancient, and it perfectly reasonable that they should employ both old and new technologies to form their bodies. But the more insightful observation is that laboratory detection of chemical potential oscil- lations is difficult and requires significant signal strength and integration time to do. Oscillations in other organ- isms might simply have not been found yet or be too weak or rapid to see easily. The diffusion constants of MinD and MinE have been measured in vivo by fluorescence correlation spectroscopy to be roughly D = 10µm2/sec78. An amplifier time of τ = 10−3 sec, a number characteris- tic an ion channel protein, gives a maximum propagation velocity of v = (D/τ )1/2 = 100 µm/sec, the speed of a fast calcium wave. For a bacterium 3µm long, this gives a round-trip transit time of 0.06 sec. Even faster speeds are possible with electrolyte ions, for which (ambipolar) diffusion constants are D ∼= 1000 µm2/sec. Acknowledgments This work was supported by the National Science Foundation under Grant No. PHY-1338376. 7 ∗ R. B. Laughlin: http://large.stanford.edu 1 C. Nusslein-Volhard, "The Identification of Genes Control- ling Development in Flies and Fishes (Nobel Lecture)," Angew. Chem. Int. Ed. Engl. 35, 2176 (1996). 2 A. Stathopoulos and D. Iber, "Studies of Morphogens: Keep Calm and Carry On," Development 140, 4119 (2013). 3 S. Sick, S. Reinker, J. Timmer, and T. Schlake, "WNT and DKK Determine Hair Follicle Spacing Through a Reaction- Diffusion Mechanism," Science 314, 1447 (2006). 4 A. Economou et al., "Periodic Stripe Formation by a Tur- ing Mechanism Operating at Growth Zones in the Mam- malian Palate," Nat. Gen. 44, 348 (2011). 5 R. Sheth et al., "Hox Genes Regulate Digit Patterning By Cntrolling the Wavelength of a Turing-Type Mechanism," Science 338, 1476 (2012). 6 P. Muller et al., "Differential Diffusivity of Nodal and Lefty Underlies a Reaction-Diffusion Patterning System," Sci- ence 336, 721 (2012). 7 T. Gregor, A. P. McGregor, and E. F. Wieschaus, "Shape and Function of the Bicoid Morphogen Gradient in Dipteran Species With Different Sized Embryos," De- vel. Biol. 316, 350 (2008). 8 A. Spirov et al., "Formation of the Bicoid Morphogen Gra- dient: An mRNA Gradient Dictates the Protein Gradi- ent," Development 136, 605 (2009). 9 H. D. Lipshitz, "Follow the mRNA: a New Model For Bi- coid Gradient Formation," Nat. Rev. Mol. Cell Bio. 10, 509 (2009). 10 O. Grimm, M. Coppey, and E. Wieschaus, "Modeling the Bicoid Gradient," Development 137 2253 (2010). 11 D. Cheung, C. Miles, M. Kreitman, and J. Ma, "Adap- tation of the Length Scale and Amplitude of the Bicoid Gradient Profile to Achieve Robust Patterning in Abnor- mally Large Drosophila melanogaster Embryos," Develop- ment 141, 124 (2014). 12 A. J. Koch and H. Meinhardt, "Biological Pattern For- mation: From Basic Mechanisms to Complex Structures," Rev. Mod. Phys. 66, 1481 (1994). 13 J. E. Pearson, "Complex Patterns in a Simple System," Science 261, 189 (1993). 21 S. Camalet, T. Duke, F. Julicher, and J. I. Prost, "Audi- tory Sensitivity Provided By Self-Tuned Critical Oscilla- tions of Hair Cells," Proc. Natl. Acad. Sci USA 97, 3183 (2000). 22 A. J. Hudspeth, F. Julicher, and P. Martin, "A Critique of the Critical Cochlea: Hopf - A Bifurcation - Is Better Than None," J. Neurophysiol. 104, 1219 (2010). 23 A. W. Scott, "The Electrophysics of a Nerve Fiber," Rev. Mod. Phys. 47, 487 (1975). 24 A. Goldbeter, "Oscillations and Waves of Cyclic AMP in Dictyostelium: A Prototype For Spatio-Temporal Organi- zation and Pulsatile Intercellular Communication," Bull. Math. Biol. 58, 1095 (2006). 25 L. F. Jaffe, "Calcium waves," Phil. Trans. Roy. Soc. B 363, 1311 (2008). 26 L. F. Jaffe, "Organization of Early Development By Cal- cium Patterns," Bioessays 21, 567 (1999). 27 M. Whitaker and J. Smith, "Introduction. Calcium Signals and Developmental Patterning," Phil. Trans. Roy. Soc. B, 363, 1307 (2008). 28 T. A. Weissman, P. A. Riqueime, L. Ivic, A. C. Flint, and A. R. Kriegstein, "Calcium Waves Propagate Through Ra- dial Glial Cells and Modulate Proliferation in the Devel- oping Neocortex,m" Neuron 43, 647 (2004). 29 E. Scemes and C. Giaume, "Astrocyte Calcium Waves: What They Are and What They Do," Glia 54, 716 (2006). 30 N. Kuga, T. Sasaki, Y. Takahara, N. Matsuki, and Y. Ikegaya, "Large-Scale Calcium Waves Traveling Through Astrocytic Networks in vivo," J. Neurosci. 31, 2607 (2011). 31 M. Junkin, Y. Lu, J. Long, P. A. Deymier, J. B. Hoying, and P. K. Wong, "Mechanically Induced Intercellular Cal- cium Communication in Confined Endothelial Structures," Biomat. 34, 2049 (2013). 32 W.-G. Choi, M. Toyota, S.-H. Kim, R. Hilleary, and S. Gilroy, "Salt Stress-Induced Ca2+ Waves Are Associ- ated With Rapid, Long-Distance Root-to-Root Signaling in Plants," Proc. Natl. Acad. Sci. 111, 5497 (2014). 33 P. Lenz and L. Sogaard-Anderson, "Temporal and Spa- tial Oscillations in Bacteria," Nat. Rev. Microbiol. 9, 565 (2011). 34 A. E. Siegman, Lasers (University Science Books, Herdon, 14 W. F. Marshall et al., "What Determines Cell Size?" BMC VA, 1986) Biol. 10, 101 (2012). 15 H. F. Nijhout, "The Control of Body Size in Insects," De- vel. Biol. 261, 1 (2013). 35 J. E. Gale and J. F. Ashmore, "An Intrinsic Frequency Limit to the Cochlear Amplifier," Nature 389, 63 (1997). 36 J. Ashmore et al., "The Remarkable Cochlear Amplifier," 16 B. Z. Stanger, "Organ Size Determination and the Limits Hearing Res. 266, 1 (2010). of Regulation," Cell Cycle 7, 318 (2008). 17 R. S. King and P. A. Newmark, "The Cell Biology of Re- generation," J. Cell. Biol. 196, 553 (2012). 18 D. W. Thompson, (1992) On Growth and Form: The Com- plete and Revised Edition (Dover, New York, 1992). 19 A. M. Turing, "The Chemical Basis of Morphogenesis," P. R. Soc. London B 237, 37 (1952). 20 T. Gold, "Hearing. II. The Physical Basis of the Action of the Cochlea," P. R. Soc. Lond B 135, 492 (1948). 37 G. A. Manley, "Evidence For an Active Process and a Cochlear Amplifier in Nonmammals," J. Neurophysiol. 86, 541 (2001). 38 B. Warren, A. N. Lukashkin, and I. J. Russell, "The Dynein-Tubulin Motor Powers Active Oscillations and Amplification in the Hearing Organ of the Mosquito," P. R. Soc. B 277, 1761 (2010). 39 N. Mhatre and D. Robert, "A Tympanal Insect Ear Ex- ploits a Critical Oscillator For Active Amplification and Tuning," Curr. Biol. 23, 1952 (2013). 40 E. C. Mora, A. Cobo-Cuan, F. Mac´ıas-Escriv´a, M. P´erez, M. Nowotny, and M. Kossl, "Mechanical Tuning of the Moth Ear: Distortion-Product Otoacoustic Emissions and Tympanal Vibrations," J. Exp. Biol. 216, 3863 (2013). 41 A. L. Hodgkin and A. F. Huxley, "A Quantitative Descrip- tion of Membrane Current and Its Application to Conduc- tion and Excitation of Nerve," J. Physiol. 117, 500 (1952). 42 R. Fitzhugh, "Mathematical Models of Excitation and Propagation in Nerve," in Biological Engineering, ed. by H. P. Schwan (McGraw Hill, New York, 1969). 43 A. L. Goldin, "Mechanisms of Sodium Channel Inactiva- tion," Curr. Opin. Neurobiol. 13, 284 (2003). 44 E. Desurvire, "Study of the Complex Atomic Susceptibility of Erbium-Doped Fiber Amplifiers," J. Lightwave Technol. 8, 1517 (1990). 45 J. S. Toll, "Causality and the Dispersion Relation: Logical Foundations," Phys. Rev. 104, 1760 (1956). 46 K. Showalter and J. J. Tyson, "Luther's 1906 Discovery of Chemical Waves," J. Chem. Ed. 64, 742 (1987). 47 D. T. Kemp, "Stimulated Acoustic Emissions From Within the Human Auditory System," J. Acoust. Soc. Am. 54, 1386 (1978). 48 A. E. Pelling, S. Dehati, E. B. Gralla, J. B. Valentine, and J. K. Gimzewski, "Local Nanomechanical Motion of the Cell Wall of Saccharomyes cerevisiae," Science 305, 1147 (2004). 49 D. M. Raskin and P. A. J. de Boer, "Rapid Pole-to-Pole Oscillations of a Protein Required For Directing Division to the Middle of Escherichia coli," Proc. Natl. Acad. Sci. USA 96, 4971 (1999). 50 G. Meacci, Min Oscillations in Escherichia Coli (VDM, Saarbruken, 2009). 51 J. R. Juarez and W. Margolin, "Changes in the Min Oscil- lation Pattern Before and After Cell Birth," J. Bacteriol. 192, 4134 (2010). 52 A. L. Marston, H. B. Thomaides, D. H. Edwards, M. E. Sharpe, and J. Errington, Polar Localization of the MinD Protein of Bacillus subtilis and Its Role in Selection of the Mid-Cell Division Site," Genes Devel. 12, 3419 (1998(). 53 M. Howard, A. D. Rutenberg, and S. De Vet, "Dynamic Compartmentalization of Bacteria: Accurate Division in E. coli," Phys. Rev. Lett. 87, 278102 (2001). 54 H. Meinhardt and P. A. J. de Boer, "Pattern Formation in Escherichia coli: A Model For the Pole-to-Pole Oscillations of Min Proteins and the Localization of the Division Site," Proc. Natl. Acad. Sci. USA 98, 14202 (2001). 55 K. C. Huang, Y. Meir, and N. S. Wingreen, "Dynamic Structures in Escherichia coli: Spontaneous Formation of MinE Rings and MinD Polar Zones," Proc. Natl. Acad. Sci. (USA) 100, 12724 (2003). 56 M. Bonny, E. Fischer-Friedrich, M. Loose, P. Schwille, and K. Kruse, "Membrane Binding of MinE Allows For a Comprehensive Description of Min-Protein Pattern For- mation," PLOS Comp. Biol. 9, e1003347 (2013). 57 J. Halatek and E. Frey, "Effective 2d Model Does Not Ac- count For Geometry Sensing By Self-Organized Proteins Patterns," Proc. Natl. Acad. Sci. USA 111, E1817 (2014). 58 H. I. Adler, W. D. Fisher, A. Cohen, and A. A. Hardigree, "Miniature Escherichia coli Cells Deficient in DNA," Proc. Natl. Acad. Sci. (USA) 57, 321 (1967). 59 P. A. J. de Boer, R. E. Crossley, and L. I. Rothfield, "A Di- vision Inhibitor and a Topological Specificity Factor Coded For by the Minicell Locus Determine Proper Placement of 8 the Division Septum in E. coli," Cell 56, 641 (1989). 60 M. Loose, E. Fischer-Friedrich, J. Ries, K. Kruse, and P. Schwille, "Spatial Regulators For Bacterial Cell Division Self-Organize into Surface Saves in Vitro," Science 320, 789 (2008). 61 V. Ivanov and K. Mizuuchi, "Multiple Modes of Intercon- verting Dynamic Pattern Formation by Bacterial Cell Di- vision Proteins," Proc. Natl. Acad. Sci. (USA) 107, 8071 (2010). 62 J. Schweizer, M. Loose, M. Bonny, K. Kruse, I Monch, and P. Schwille, "Geometry Sensing By Self-Organized Protein Patterns," Proc. Natl. Acad. Sci. (USA) 109, 15283 (2012). 63 Z. Hu, E. P. Gogol, and J. Lutkenhaus, Dynamic Assembly of MinD on Phospholipid Vesicles Regulated by ATP and MinE," Proc. Natl. Acad. Sci. (USA) 99, 6761 (2002). 64 L. L. Lackner, D. M. Raskin, and P. A. J. de Boer, "ATP- Dependent Interactions Between Escherichia coli Min Pro- teins and the Phospholipid Membrane Pin vitro," J. Bac- teriol. 185, 735 (2003). 65 K. D. Young, "Bacterial Shape: Two-Dimensional Ques- tions and Possibilities," Annu. Rev. Microbiol. 64, 223 (2010). 66 K. C. Huang, D. W. Ehrhardt, and J. W. Shaevitz, "The Molecular Origins of Chiral Growth in Walled Cells," Curr. Opin. Microbiol. 15, 1 (2012). 67 W. R. DiLuzio, L. Turner, M. Mayer, P. Garstecki, D. B. Weibel, H. C. Berg, and G. M. Whitesides, "Escherichi coli Swim on the Right-Hand Side," Nature 435, 1271 (2005). 68 E. Lauga, W. R. DiLuzio, G. M. Whitesides, and M. A. Stone, "Swimming in Circles: Motion of Bacteria Near Solid Boundaries," Biophys. J. 90, 400 (2006). 69 E. C. Garner, R. Bernhard, W. Wang, X. Zhuang, D. Z. Rudne, and T Mitchison, "Coupled, Circumferential Mo- tions of the Cell Wall Synthesis Machinery and MreB Fil- aments in B. subtilis," Science 333, 222 (2011). 70 J. Dom´ıinguez-Escobar, A. Chastanet, A. H. Crevenna, V. Fromion, R. Wedlich-Soldner, and R Carballido-L´opez, "Precessive Movement of MreB-Associated Complexes in Bacteria," Science 333, 225 (2011). 71 S. van Teeffelen, S. Wang, L. Furchgott, K. C. Huang, N. S. Wingreen, J. W. Shaevitz, and Z. Gital, "The Bacterial Actin MreB Rotates, and Rotation Depends on Cell-Wall Assembly," Proc. Natl. Acad. Sci. 108, 15882 (2011). 72 J. Hazel et al., "Changes in Cytoplasmic Volume Are Suf- ficient to Drive Spindle Scaling," Science 342, 853 (2013). 73 M. C. Good, M. D. Vahey, A. Skandarjah, D. A. Fletcher, and R. Heald, "Cytoplasmic Volume Modulates Spindle Size During Embryogenesis," Science 324, 856 (2013). 74 L. Yang and I. R. Epstein, "Oscillatory Turing Patterns in Reaction-Diffusion Systems With Two Coupled Layers," Phys. Rev. Lett. 90, 178303 (2003). 75 H. Meinhardt H, "Models of Biological Pattern Formation: From Elementary Steps to the Organization of Embryonic Axes," Curr. Top. Dev. Biol. 81, 1 (2008). 76 S. Kondo and T. Miura, "Reaction-Diffusion Model as a Framework for Understanding Biological Pattern Forma- tion," Science 329, 1616 (2010). 77 D. M. Umulis and H. G. Othmer, "Mechanisms of Scaling in Pattern Formation," Development 140, 4830 (2013). 78 G. Meacci, J. Ries, E. Fischer-Friedrich, N. Kahya, P. Schwille, and K. Kruse, "Mobility of Min-proteins in Es- cherichia coli Measured by Fluorescence Correlation Spec- troscopy," Phys. Biol. 3, 255 (2006).
1910.14237
2
1910
2019-12-17T23:39:18
Polarisable force fields: What do they add in biomolecular simulations?
[ "physics.bio-ph", "q-bio.BM" ]
The quality of biomolecular simulations critically depends on the accuracy of the force field used to calculate the potential energy of the molecular configurations. Currently, most simulations employ non-polarisable force fields, which describe electrostatic interactions as the sum of Coulombic interactions between fixed atomic charges. Polarization of these charge distributions is incorporated only in a mean-field manner. In the past decade, extensive efforts have been devoted to developing simple, efficient, and yet generally applicable polarisable force fields for biomolecular simulations. In this review, we summarise the latest developments in accounting for key biomolecular interactions with polarisable force fields and applications to address challenging biological questions. In the end, we provide an outlook for future development in polarisable force fields.
physics.bio-ph
physics
Polarisable force fields: What do they add in biomolecular simulations? V. S. Sandeep Inakollu,​1,2,3​ Daan P. Geerke,​4,*​ Christopher N. Rowley,​5,*​ Haibo Yu​1,2,3,* 1. School of Chemistry and Molecular Bioscience, University of Wollongong, Wollongong NSW 2522, Australia 2. Molecular Horizons, University of Wollongong, Wollongong NSW 2522 Australia 3. Illawarra Health and Medical Research Institute, Wollongong, New South Wales, 2522, Australia 4. AIMMS Division of Molecular and Computational Toxicology, Department of Chemistry and Pharmaceutical Sciences, Vrije Universiteit Amsterdam, De Boelelaan 1108, 1081 HZ Amsterdam, the Netherlands 5. Department of Chemistry, Memorial University of Newfoundland, St. John's, Newfoundland and Labrador, Canada *Corresponding authors: ​[email protected]​, ​[email protected]​, ​[email protected] 1 Abstract The quality of biomolecular simulations critically depends on the accuracy of the force field used to calculate the potential energy of the molecular configurations. Currently, most simulations employ non-polarisable force fields, which describe electrostatic interactions as the sum of Coulombic interactions between fixed atomic charges. Polarisation of these charge distributions is incorporated only in a mean-field manner. In the past decade, extensive efforts have been devoted to developing simple, efficient, and yet generally applicable polarisable force fields for biomolecular simulations. In this review, we summarise the latest developments in accounting for key biomolecular interactions with polarisable force fields and applications to address challenging biological questions. In the end, we provide an outlook for future development in polarisable force fields. Introduction Atomistic modelling plays an increasingly important role in understanding the structure-function-dynamics relationship in biomolecular systems. This understanding now facilitates various types of molecular engineering that would have been impossible without the insights provided by modelling ​[1]​. The accuracy and predictive power of molecular dynamics (MD) simulations based on all-atom force fields are steadily improving due to the parallel improvements in high-performance computing hardware, more accurate methods for calculating the potential energy of a conformation, and more efficient methods for conformational sampling. Nowadays, µs-length simulations of systems containing hundreds of thousands of atoms are performed routinely. With specialised supercomputers, it has been 2 possible to perform millisecond-length simulations,​[2] and the simulations of entire cellular structures have been attempted.​[3] The general form of the widely-used conventional force fields dates back to the pioneering work by Lifson's group.​[4,5] It consists of the bonded interactions (bonds, valence angles, dihedral angles) and the nonbonded interactions (both electrostatic and van der Waals). The van der Waals term is often described by a Lennard-Jones form, and the electrostatic interactions are described using Coulomb's law, with fixed partial charges preassigned to each atom according to the adopted force field. This type of force field is called an additive or non-polarisable force field. Force field developers have a variety of strategies to parameterise the partial charges.​[6,7] One common feature among them is that the polarisation effect is in which the partial charges and dipole moments are treated in a mean-field manner, enhanced compared to their gas-phase values, mimicking the effect of induced polarisation in an average way. Although this model is simple, they have benefited from almost 40 years of parameterization refinements, and they have provided a wealth of information into complex molecular systems.​[1] The inherent limitation of these models is that they are incapable of describing the change of polarisation of molecules when they adopt different conformations or encounter different interacting partners over the course of a simulation. For example, the polarisation of a solute is expected to increase when it moves from a non-polar region of the system into a polar region, but this effect is neglected by conventional non-polarisable models. Developing computational models that account induced polarisation has been a longstanding objective in computational biophysics.​[8] However, the broad adoption of for 3 polarisable force fields in biomolecular simulations was hampered by the limited availability of model parameters and the increased computational cost. In recent years, there have been sustained efforts by several groups towards devising and parameterizing polarisable force fields for biomacromolecules. At the same time, the development of high-performance computing has allowed sufficient conformational sampling of systems of biological interest using these polarisable models​[9,10]​. For instance, using NAMD, the computational effort required for an MD simulation using a polarisable model is roughly double that of the non-polarisable counterpart, making these simulations tractable if sufficient computing resources are available​[10]​. There are at least three different methods to account for explicit polarisation in classical force fields:​[11] the Point-Polarisable Dipole (PPD)​[12,13]​, Fluctuation Charge (FQ)​[14,15] and Drude Oscillator (DO) ​[16] (or called Shell Model​[17]​, and Charge-on-Spring model​[18]​). Combined models can be found in the literature too. Huang et al. recently demonstrated that it is possible to map the electrostatic model optimised in the Drude force field onto the multipole and induced dipole model and illustrated the equivalency between DO and PPD.​[19] This review article will focus on the latest developments in and applications of polarisable force fields for biomolecular simulations. We will not add extensive general references to various polarisable models, and readers are referred to the latest review articles.​[20 -- 22] First, we briefly review the recent development in dealing with challenging molecular interactions and highlight some of the latest applications of polarisable force fields. Finally, we present a summary and outlook. 4 Fundamental key interactions Additive force fields are the most commonly used force fields in biomolecular simulations. However, their accuracy can be limited by their use of fixed atomic charges. This is particularly significant for modelling processes where electrostatic interactions are changing and fluctuating or where induced polarisation is an essential part of the interactions. Compared to additive models, explicitly accounting for polarisation can increase the force field parameter sets in terms of transferability of their accuracy to describe intermolecular interactions in environments of different polarities. ​[23] Consequently, it is challenging to describe some key biomolecular interactions using additive models, such as cation -- 𝜋 and metal/molecular -- ion interactions. As described below, recent efforts have focused on developing polarisable force fields to describe such interactions accurately (Figure 1)​. Moreover, the deficiencies in the models currently used to describe London dispersion interactions are noted. 5 Figure 1:​ (a) Cation-π interactions in the biomolecular system illustrated by a cation-π interaction between Lys1 and Trp10 in the HP peptide (PDB ID: 2EVQ). The difference in electron density distributions between the interacting and non-interacting states (left) shows that the Trp π-electron density is polarised towards the cationic NH​3​ lysine (blue) away from the atomic nuclei (red). In the CHARMM-Drude model (right) this type of cation-π interaction is approximated by Drude oscillators tethered to the non-hydrogen atoms and an additional charge at the centre of the π ring (black point). (b) Metal and molecular ion interactions illustrated by Z-DNA crystal with 2 Mg​2+​ (PDB ID: 1LJX). The CHARMM-Drude model (right) accurately describes the interactions between Mg​2+​ and phosphate groups of the nucleic acid (taken from Ref. ​[24]​). (c) Other biologically relevant elements and functional groups illustrated by covalent-modifier ibrutinib bound to TgCDPK1 (PDB ID: 4IFG, taken from Ref.​[25]​). The electron densities of water molecules coordinated to a model thiolate are polarised by the anionic charge (left). +​ group of the Cation-​𝜋​ and ​𝜋-𝜋​ interactions Cation-𝜋 interactions commonly occur between the positively charged cations and negatively charged 𝜋 electron-rich cloud in the aromatic ring in the charged and aromatic amino acid or nucleic acids.​[26] These interactions are highly anisotropic in nature. The polarisation and the charge redistributions are essential to model these interactions correctly. Rupakheti et al.​[27] studied the commonly occurring cation-𝜋 interactions in the proteins between the aromatic and charged amino acids, by comparing the potentials of mean force (PMF) for a series of prototypical cation-𝜋 models with both CHARMM36 (C36) and the Drude-2013 polarisable force field.​[28] Based on the reversible association PMFs, accounting for polarisation globally enhanced the description of the cation-𝜋 interactions. they showed that explicitly They also noted the challenges in accurately describing the interactions responsible for amino acid cation-𝜋 interactions. Lin and MacKerell​[29] systematically optimised the CHARMM Drude-2013 polarisable force field parameters​[28] for cation-𝜋 and anion-aromatic ring interactions, targeting the QM interaction energies and geometries. The atom pair-specific Lennard-Jones parameters along with virtual particles as selected ring centroids were 6 introduced. The refined CHARMM Drude-2013 protein force field has been shown to in reproducing the ion-π pair distances observed in provide a significant experimental protein structures ​(Figure 2)​. Zhang et al.​[30] developed the AMOEBA polarisable force field for aromatic molecules and nucleobases, in which their parameters improvement were parameterised against the properties in the gas phase with QM calculations and experimental values in the condensed phase. They further extend the development to a full set of AMOEBA force fields for nucleic acids.​[31] Figure 2:​ The cation-π interaction in the HP peptide (PDB ID: 2EVQ). (a) Structure of the HP peptide with the analysed Lys1-Trp10 cation-π pair, where oxygen is in red, nitrogen in blue, carbon in white, and water molecules are not shown. (b) Normalised distribution of the distances between the 6-membered ring centre of Trp10 and the side-chain nitrogen of Lys1 computed from simulations with Drude-2013 (black) and with Drude-2013-CP (red) compared to those calculated from NMR structures (Exp, blue). Reproduced from ​Ref. [29]​. Metal and molecular ion interactions 7 Metal ions are fundamental to the structure and function of many biological systems, where they may interact with solvent, proteins, membranes and nucleic acids. The presence of the metal ion strongly alters the local electrostatic environment. Several studies have pointed out the intrinsic limitations of additive force fields in studying metal ion interactions.​[32,33] Parameters have been developed for the set of biologically relevant ions for both the Drude and AMOEBA force fields.​[34,35] The AMOEBA force field was used to study the selectivity for Ca​2+ and Mg​2+ ions for various protein binding pockets. It was shown that unless polarisation was included, the smaller ion Mg​2+ is always favoured over the larger ion Ca​2+​.​[36] Another notable recent development includes polarisable models for biologically relevant molecular ions.​[37] For instance, phosphate groups are essential components of nucleic acids. Their interactions with the surrounding solvents, metal ions, and proteins facilitate the binding and folding motions in the nucleic acids. Lemkul and MacKerell ​[38] ​[24] studied the interactions of phosphate analogues, including dimethyl and Villa et al. phosphate (DMP) and methyl phosphate (MP), with the Mg​2+ ion with the Drude polarisable force field. The Mg​2+​-phosphate-binding free energies calculated using the Drude model have better agreement with the QM and experimental data. Furthermore, the refined complete set of Drude polarisable force field for DNA and RNA has been reported and validated.​[39 -- 41] Similar work has been carried out for the AMOEBA force field.​[42] While these models provide potential energy surfaces that are in reasonable agreement with QM results, energy decomposition analysis (EDA) has revealed that the relative magnitude of the components of the interaction energy of the polarisable MM and QM models can be very different. In this analysis, the charge-penetration (CP), charge-transfer (CT), dispersion, permanent electrostatic, and polarisation interactions in water -- water, water -- ion, and ion -- protein model compounds were calculated using EDA of the DFT 8 interaction energy with the absolutely localised molecular orbitals (ALMO) scheme and compared to the components of the AMOEBA interaction energy.​[43,44] AMOEBA does not include CP and CT terms, but in water -- water interactions, the 14-7 potential used to represent van der Waals interactions in the AMOEBA model partially compensated for these effects. This cancelation of error was less effective for water -- halide, water -- divalent cation, and Ca​2+​-protein models, where the magnitudes of permanent electrostatic and polarisation interactions in the AMOEBA model deviated significantly from the EDA results. These studies serve to guide the future parametrisation of explicit functional forms for short-range contributions from CP and/or CT.​[45 -- 47] Other biologically important elements and groups Cysteine is a unique sulphur amino acid involved in various biological processes, including protein-ligand binding, catalytic reactions, and post-translational modifications. Due to the presence of the thiol group, which has a moderate pKa, cysteine can exist in its anionic form under physiological conditions. Non-polarisable force fields have limited success in describing the structure and hydration energies of these highly polarisable ions. Lin et al.'s development of a CHARMM-Drude model for polyatomic ions provided the first polarisable model for thiolates.​[37] Williams and Rowley​[48] showed that the Drude polarisable model predicted the structural and energetic properties of methylthiolate in good agreement with QM/MM MD simulations, while the conventional MM model overestimated its solvation free energy. Recently, Drude polarisable force field parameters have been developed for halogen-containing compounds, which will allow this model to be used to model the binding of halogenated drugs to protein targets.​[49] 9 van der Waals interactions Although these polarisable models account for the induction of an atomic dipole from the electric field created by the environment around the atom, the instantaneous-dipole -- induced dipoles that give rise to the London dispersion interactions are not captured. The pairwise Lennard-Jones potential or a similar 14-7 potential has been adopted to account for Pauli repulsive and dispersion forces in the polarisable force fields. As the electrostatic components of these force fields have changed, the van der Waals parameters of conventional force fields are no longer appropriate, so new parameters have to be determined for use with the polarisable force fields. Typically, non-bonded parameters of polarisable models are still assigned empirically based on bulk physical properties of liquids. While polarisable force fields typically have static charges and dipole-moments that are closer to their gas-phase QM estimates than additive force fields, molecular dispersion C​6 parameters are typically too high. ​[50,51]​. Recently, new methods have been developed to define dispersion parameters from quantum chemical calculations, which has the potential to simplify force field development and make the models more transferable.​[52,53] Protein simulations Protein structure and dynamics are other areas where induced polarisation is expected to have a significant effect. For example, when proteins fold to form α-helices, the NH and C=O moieties of the amide backbone form strong hydrogen bonds. The polarisation of these bonds results in a cooperative effect, where the strength of the hydrogen bonds increases as the number of turns in the helix increases.​[54] Likewise, the cooperativity of hydrogen bonds between polar side chains can stabilize the folded state of a protein. The accurate description 10 of the relative stability and transition rates between unfolded/misfolded and folded states will likely require explicit treatment of induced polarisation.​[55] These issues are particularly relevant in the simulation of intrinsically disordered proteins IDPs are involved in several pathological disorders, (IDP). including cancer and neurodegenerative disorders​[56]​. IDPs are characterised by the lack of well-defined tertiary structure. Instead, they exist in an interconverting ensemble of conformations. The amino acid sequence in IDPs is enriched with polar and charged amino acids, and have relatively low numbers of hydrophobic amino acids, which are essential for protein core formation.​[57] Both Amber and CHARMM additive force fields have recently been refined to provide a better description of IDPs, although their performance is inconsistent.​[58,59] Treatment of explicit polarisation may be needed to model the diverse range of structure IDPs exist in.​[57,60] Wang et al.​[61] conducted a study to compare the performance of non-polarisable and polarisable force fields for protein structural refinement, protein folding, and simulating IDPs. They showed that the inclusion of explicit polarisation improves accuracy in protein structure refinement and the description of IDP conformational ensembles. This study also noted the difficulties for the polarisable force field to sample the native structures in the selected proteins. To address this limitation, future work is required to further refine the parameters. This may well comprise improving the description of dispersion, which was recently shown to be important for the simulation of IDPs.​[58] Water dynamics on the surface of proteins play a significant role in protein folding and unfolding. Ngo et. al.​[62] studied the hydration free energies of amino acid side chains, protein-water and protein-protein interactions, and the hydrogen-bond lifetime with the CHARMM additive C36 and Drude polarisable force fields. The side chain hydration 11 energies predicted by the CHARMM Drude force field are generally in better agreement with the experimental data than that of the C36 force field, except for the acidic amino acid side chains. The development of revised CHARMM-Drude parameters for molecular ions may help resolve this issue.​[37] In the simulations with the CHARMM Drude force field, stronger interactions and longer-lived hydrogen bonds between the first hydration shell and the protein were observed. Furthermore, the first solvation shell prevents other waters from accessing the protein surface. Hazel et al.​[63] studied the folding free energy landscapes of C-terminal β-hairpin of the B1 domain of streptococcal protein G (GB1) using replica exchange umbrella sampling simulations with two non-polarisable force fields (C36 and C22*) and the CHARMM-Drude-2013 polarisable force field. Surprisingly, the C22* and CHARMM-Drude model agreed better with the experimental studies of GB1 folding, while C36 over stabilises the β-hairpin. Current literature suggests that more validation studies and continuous refinement of the polarisable force fields are needed for it to be widely applicable in simulating protein dynamics. Protein-ligand interactions Electrostatic interactions can play a major role in protein​ -- ​ligand and enzyme​ -- ​substrate interactions. Often the protein binding sites and the enzyme active sites encompass a heterogeneous environment that can also include water molecules and metal ions. This presents challenges for additive force fields, particularly for highly-charged species. Qi et al. used the AMOEBA polarisable force field in designing inhibitors for fructose-bisphosphate aldolase A (ALDOA).​[64] ALDOA converts fructose-1,6 bisphosphate (FDP) into 12 glyceraldehyde-3-phosphate (GAP) and dihydroxyacetone phosphate. Substrate-mimicking inhibitors for ALDOA are typically highly charged. The AMOEBA simulations were applied to model the binding of a series of naphthalene-2,6-diyl bisphosphate analogues and rank their relative binding free energies, which match experimental data well. Panel et al. ​[65] studied binding specificity between the PDZ domain and C-terminal peptides of its target proteins, which form the building blocks of eukaryotic signalling pathways. It was found that the additive force field AMBER ff99SB over-stabilises salt-bridge interactions and the Drude force field significantly reduced errors for those involving ionic mutations. This suggests that electronic polarisation can be crucial to describe ionic interactions in buried regions. Welborn and Head-Gordon​[66] used the AMOEBA force field to study the electric field-driven enzyme catalytic reaction in the enzyme ketosteroid isomerase (KSI). The calculated electric fields induced by the active site of KSI on the carbonyl probe in 19-NT ligand are -108±4.9 MV/cm with AMOEBA. The authors also showed that simulations without mutual polarisation reduced the electric field to −68.08 ± 3.1 MV/cm. The encouraging agreement with the experimental value (i.e., 120-150 MV/cm) for AMOEBA simulations highlights the need for explicit polarisation to capture the changes of the electric fields at the enzyme active site. Another area of interest is the O​2 binding and diffusion in biomolecular systems. O​2 is a neutral but highly polarisable molecule and non-polarisable force fields represent its interactions with the environment with van der Waals interactions only.​[67] Torabifard and Cisneros compared O​2 diffusion in Alk with the AMBER and AMOEBA force fields.​[68] The PMF based on both force fields consistently showed a passive transport of O​2 from the 13 surface of the protein to the active site. However, the PMF by AMOEBA shows a larger barrier for diffusion of the co-substrate out of the active site than the non-polarisable force field. It has been suggested that explicit polarisation is crucial to adequately describe the interactions between O​2​ (neutral albeit highly polarisable) and its environment. Ion channels Electrostatics and polarisation also play an important role in the mechanisms of ion channel gating and conduction.​[69] Peng et al. showed that they were able to reproduce the experimental conductance in Gramicidin A with the AMOEBA force field.​[70] Sun and Gong​[71] modelled the transition in the voltage-gated sodium channel (Na​V​) from its resting state to the pre-active state using the CHARMM-Drude force field. They were able to show the conformational changes of Na​V from the resting state to the pre-active state. The polarisation of the 𝜋-electrons in Phe56 by the positively charged Arg3 in Na​V was found to stabilise the protein structure when the charged gating residues pass the hydrophobic constriction site during activation. Polarisable force fields have been used to study other ion channels as well.​[72,73] Membrane permeation Biological membranes are composed of a bilayer of mixed lipid components with membrane proteins embedded in them. Many cellular signalling and metabolic processes require selective passage of ions or small molecules across the membrane either through non-facilitated permeation through the lipid bilayer or by facilitation by membrane-spanning proteins. These structures inherently possess various electrostatic environments, as ionic or 14 polar headgroups face the interior and exterior solutions to form a water​ -- ​membrane interface while the interior of the membrane is composed of non-polar saturated and unsaturated lipid tails. As a consequence, molecules permeating through the membrane experience different degrees of polarisation depending on their positions in the membrane. Induced polarisation can play a significant role in non-facilitated membrane permeation. Small molecules permeating a lipid bilayer cross between the polar aqueous solution, through the ionic water​ -- ​bilayer interface, and through the non-polar lipid tails in the interior of the bilayer. This range of electrostatic environments results in large shifts in the induced polarisation of permeating solutes. Riahi and Rowley explored these effects in simulations of the permeation of water and hydrogen sulphide through a DPPC lipid bilayer using the CHARMM-Drude polarisable force field.​[74] The dipole moment of the permeating water molecule was largest ( <μ>=2.5 D) in the aqueous phase where there are strongly-polarising hydrogen bonds with other water molecules. This polarisation decreases as the water molecules enter the bilayer, reaching a minimum at the centre of the membrane, where the dipole moment is ~1.9 D. Hydrogen sulphide shows a similar but less pronounced trend, where the average solute dipole decreases from 1.2 D to 1.0 D ​(Figure 3)​. This highlights an apparent paradox in the induced polarisation of solutes in condensed phases; highly polarisable molecules such as hydrogen sulphide experience a smaller degree of induced polarisation than the less polarisable water molecules. This reflects that the atomic radii of atoms also increase with their polarisability, so highly polarisable atoms, like S and C, may well be too large to participate in strong, short-range electrostatic interactions that result in a strong induced polarisation effect. 15 a and Figure 3. The average dipole moment of a water molecule hydrogen (blue) sulphide (yellow) permeating through a DPPC lipid bilayer, the represented CHARMM-Drude polarisable force field. Adapted from Ref. [74]​. using QM/MM simulations and computational vibrational spectroscopy QM/MM MD simulations are powerful methods to study how the environment affects the reactivity or spectroscopic properties of a critical component. An immediate concern is that the enhanced partial charges in additive force fields will create an inconsistent and unbalanced description of the interactions between the QM part and the MM part in combined QM/MM simulations. Polarisable force fields may offer a solution to this issue, and there have been many reports where a QM/MM model was constructed using a polarisable MM model.​[75,76] The accuracies of these simulations depend on the QM model, the MM model, and the interactions between QM and MM. König ​et al. systematically studied the hydration free energies of 12 small molecules with QM/MM simulations with the CHARMM force field and the CHARMM-Drude polarisable force field.​[77] Despite the 16 potential for the polarisable model to provide more accurate results, the resulting QM/MM hydration free energies were inferior to purely classical results, with the QM/MM(Drude) predictions being only marginally better than the QM/MM(non-polarisable) results. Ganguly et al.​[78] reported the first systematic assessment of a polarisable force field in QM/MM studies of enzymatic reactions. In the cases of the Claisen rearrangement in chorismate mutase and the hydroxylation reaction in p-hydroxybenzoate hydrolase, the authors observed that explicit MM polarisation has moderate effects on activation and reaction (free) energies. They concluded that further validation work is required to establish the best QM/MM-based procedure for handling polarisation effects in enzymatic reactions. Polarisable force fields have also been applied to the prediction of vibrational spectra, especially where the vibrational models are highly anharmonic in nature or are sensitive to the surrounding electrostatic environment​[79]​. Semrouni et al.​[80] and Thaunay et al.​[81,82] and their applied the AMOEBA force vibrational calculate spectra field to temperature-dependence using the Fourier transform of the dipole autocorrelation function. Explicit polarisation could provide improved sensitivity of the spectra to the environment by rigorously including solvent​ -- ​solute interactions like hydrogen bonds. Furthermore, combined QM and polarisable force field simulations are an attractive method to predict and understand the infrared spectra of molecules in solution and a biomolecular system. ​[83] Conclusions and Outlook In the past decades, we have witnessed impressive progress in the development of polarisable force fields and their application in biomolecular simulations. This has been enabled by efficient software development and continuous refinement of force field parameters. The 17 applications have provided many new insights into biological processes, where explicit polarisation is crucial. At the same time, more systematic validation is needed to understand and improve some of the limitations in the current models, including both the underlying physical models and their parameterisation. The development of automated and systematic parameterisation techniques is particularly promising. Acknowledgements The authors thank Dr Jing Huang (Westlake University, China), Dr Qiantao Wang (Sichuan University, China) and Mr Koen Visscher (Vrije Universiteit Amsterdam, the Netherlands) for comments on this article. We wish to acknowledge the Australian Government for an Australian International Postgraduate Award scholarship for V.I. DPG acknowledges financial support from the Netherlands Organization for Scientific Research (NWO, VIDI grant 723.012.105). CNR thanks NSERC of Canada for funding through the Discovery Grants program (RGPIN-05795-2016). This research was in part supported under the Australian Research Council's Discovery Project (DP170101773, H.Y.) and National Health and Medical Research Council Project Grant (APP1145760, H.Y.). This article is dedicated to the memory of Prof Herman J.C. Berendsen, one of the founding fathers of biomolecular simulations. References and recommended readings Papers of particular interest, published within the period of review, have been highlighted as: · of special interest [9] ​Tinker-HP, a massively parallel molecular dynamics package, has been developed and it enables efficient molecular dynamics simulations with the AMOEBA force field. 18 [29] The cation-𝜋 and anion-ring interactions in the CHARMM Drude-2013 force field have been refined. The refined force field leads to a significant improvement in reproducing the geometric properties of the ion-𝜋 interactions in experimental protein structures. [53] A protocol was proposed to parameterise a polarisable force field based on quantum mechanical calculations. It showed that combining C​6 and C​8 attractive terms together with a C​11 repulsive potential yields satisfying models and this demonstrated that explicit inclusion of higher-order dispersion terms could be viable in polarisable force field development. [78] The effects of explicit MM polarisation in QM/MM simulations of enzymatic reactions have been systematically studied in two systems. The authors concluded that further validation is required to establish the best QM/MM-based procedure for dealing with the effects of polarisation in enzymatic reactions. ·· of outstanding interest [61] Long-time-scale MD simulations were carried out to evaluate the effect of explicit polarisation in protein simulations. The authors showed that polarisable force fields have better performance in protein structure refinement and sample conformational ensembles of IDPs. They also noted the limitation of the current polarisable force fields in predicting the protein native state. [63] This study compared the folding free energy landscape of GB1 with non-polarisable and polarisable CHARMM force fields. The additive CHARMM36 force field was found to overstablize the folded hairpin structure of GB1. The CHARMM 19 Drude-2013 force field predicted the relative stability more accurately, but several areas where this force field could be improved further were also identified. Additionally, they showed that tryptophan fluorescence is insufficient for capturing the full β-hairpin folding pathway. [66] The authors reported the effects of conformational dynamics on the fluctuations of electric fields on the carbonyl probe in ketosteroid isomerase (KSI) with the AMOEBA force field. It was found that Asp103 promotes large fluctuations in the electric field along the functional cycle of KSI. 5. 1. van Gunsteren WF, Bakowies D, Baron R, Chandrasekhar I, Christen M, Daura X, Gee P, Geerke DP, Glättli A, Hünenberger PH, et al.: ​Biomolecular Modeling: Goals, Problems, Perspectives​. ​Angew Chem Int Ed​ 2006, ​45​:4064 -- 4092. 2. Lindorff-Larsen K, Piana S, Dror RO, Shaw DE: ​How Fast-Folding Proteins Fold​. Science​ 2011, ​334​:517 -- 520. Feig M, Yu I, Wang P, Nawrocki G, Sugita Y: ​Crowding in Cellular Environments at 3. an Atomistic Level from Computer Simulations​. ​J Phys Chem B​ 2017, 121​:8009 -- 8025. 4. Lifson S, Warshel A: ​Consistent Force Field for Calculations of Conformations, Vibrational Spectra, and Enthalpies of Cycloalkane and n-Alkane Molecules​. ​J Chem Phys​ 1968, ​49​:5116 -- 5129. van Gunsteren WF: ​The Roots of Bio-Molecular Simulation: The Eight-Week CECAM Workshop 'Models for Protein Dynamics' of 1976​. ​Helv Chim Acta​ 2019, 102​. 6. Bayly CI, Cieplak P, Cornell W, Kollman PA: ​A well-behaved electrostatic potential based method using charge restraints for deriving atomic charges: the RESP model​. ​J Phys Chem​ 1993, ​97​:10269 -- 10280. on the free enthalpy of hydration and solvation: The GROMOS force-field parameter sets 53A5 and 53A6​. ​J Comput Chem​ 2004, ​25​:1656 -- 1676. 8. Warshel A, Kato M, Pisliakov AV: ​Polarizable Force Fields: History, Test Cases, and Prospects​. ​J Chem Theory Comput​ 2007, ​3​:2034 -- 2045. 9. Lagardère L, Jolly L-H, Lipparini F, Aviat F, Stamm B, Jing ZF, Harger M, Torabifard H, Cisneros GA, Schnieders MJ, et al.: ​Tinker-HP: a massively parallel molecular dynamics package for multiscale simulations of large complex systems with advanced point dipole polarizable force fields​. ​Chem Sci​ 2018, ​9​:956 -- 972. High-Performance Scalable Molecular Dynamics Simulations of a Polarizable Force Field Based on Classical Drude Oscillators in NAMD​. ​J Phys Chem Lett​ 2011, 2​:87 -- 92. 7. Oostenbrink C, Villa A, Mark AE, Gunsteren WFV: ​A biomolecular force field based 10. Jiang W, Hardy DJ, Phillips JC, MacKerell AD, Schulten K, Roux B: 11. Yu H, van Gunsteren WF: ​Accounting for polarization in molecular simulation​. 20 13. Van Belle D, Couplet I, Prevost M, Wodak SJ: ​Calculations of electrostatic properties 19. Huang J, Simmonett AC, Pickard FC, MacKerell AD, Brooks BR: ​Mapping the Drude Comput Phys Commun​ 2005, ​172​:69 -- 85. 12. Warshel A, Levitt M: ​Theoretical studies of enzymic reactions: Dielectric, electrostatic and steric stabilization of the carbonium ion in the reaction of lysozyme​. ​J Mol Biol​ 1976, ​103​:227 -- 249. in proteins: Analysis of contributions from induced protein dipoles​. ​J Mol Biol 1987, ​198​:721 -- 735. J Phys Chem​ 1991, ​95​:3358 -- 3363. Application to liquid water​. ​J Chem Phys​ 1994, ​101​:6141 -- 6156. 14. Rappe AK, Goddard WA: ​Charge equilibration for molecular dynamics simulations​. 15. Rick SW, Stuart SJ, Berne BJ: ​Dynamical fluctuating charge force fields: 16. Drude P: ​The theory of optics​. Dover Publications; 1959. 17. Born M, Huang K: ​Dynamical Theory of Crystal Lattices​. Clarendon Press; 1988. 18. Straatsma TP, McCammon JA: ​Molecular Dynamics Simulations with Interaction Potentials Including Polarization Development of a Noniterative Method and Application to Water​. ​Mol Simul​ 1990, ​5​:181 -- 192. polarizable force field onto a multipole and induced dipole model​. ​J Chem Phys 2017, ​147​:161702. 20. Lemkul JA, Huang J, Roux B, MacKerell AD: ​An Empirical Polarizable Force Field Based on the Classical Drude Oscillator Model: Development History and Recent Applications​. ​Chem Rev​ 2016, ​116​:4983 -- 5013. 21. Jing Z, Liu C, Cheng SY, Qi R, Walker BD, Piquemal J-P, Ren P: ​Polarizable Force Fields for Biomolecular Simulations: Recent Advances and Applications​. ​Annu Rev Biophys​ 2019, ​48​:371 -- 394. 22. Melcr J, Piquemal J-P: ​Accurate biomolecular simulations account for electronic polarization​. 2019, doi:arXiv:1909.03732. 23. Geerke DP, van Gunsteren WF: ​Calculation of the Free Energy of Polarization: Quantifying the Effect of Explicitly Treating Electronic Polarization on the Transferability of Force-Field Parameters​. ​J Phys Chem B​ 2007, ​111​:6425 -- 6436. 24. Villa F, MacKerell AD, Roux B, Simonson T: ​Classical Drude Polarizable Force Field Model for Methyl Phosphate and Its Interactions with Mg ​2+​. ​J Phys Chem A 2018, ​122​:6147 -- 6155. 25. Awoonor-Williams E, Walsh AG, Rowley CN: ​Modeling covalent-modifier drugs​. Biochim Biophys Acta BBA - Proteins Proteomics​ 2017, ​1865​:1664 -- 1675. 26. Gallivan JP, Dougherty DA: ​Cation-π interactions in structural biology​. ​Proc Natl Acad Sci​ 1999, ​96​:9459 -- 9464. 27. Rupakheti CR, Roux B, Dehez F, Chipot C: ​Modeling induction phenomena in amino acid cation-pi interactions​. ​Theor Chem Acc​ 2018, ​137​:174. 28. Lopes PEM, Huang J, Shim J, Luo Y, Li H, Roux B, MacKerell AD: ​Polarizable Force Field for Peptides and Proteins Based on the Classical Drude Oscillator​. ​J Chem Theory Comput​ 2013, ​9​:5430 -- 5449. Interactions Using the Drude Polarizable Empirical Force Field for Proteins​. ​J Comput Chem​ 2019, doi:10.1002/jcc.26067. 30. Zhang C, Bell D, Harger M, Ren P: ​Polarizable Multipole-Based Force Field for Aromatic Molecules and Nucleobases​. ​J Chem Theory Comput​ 2017, ​13​:666 -- 678. 31. Zhang C, Lu C, Jing Z, Wu C, Piquemal J-P, Ponder JW, Ren P: ​AMOEBA 29. Lin F, MacKerell AD: ​Improved Modeling of Cation-π and Anion-Ring 21 38. Lemkul JA, MacKerell AD: ​Balancing the Interactions of Mg ​2+​ in Aqueous Solution Polarizable Atomic Multipole Force Field for Nucleic Acids​. ​J Chem Theory Comput 2018, ​14​:2084 -- 2108. 32. Yu H, Mazzanti CL, Whitfield TW, Koeppe RE, Andersen OS, Roux B: ​A Combined Experimental and Theoretical Study of Ion Solvation in Liquid ​N -Methylacetamide​. ​J Am Chem Soc​ 2010, ​132​:10847 -- 10856. 33. Li H, Ngo V, Da Silva MC, Salahub DR, Callahan K, Roux B, Noskov SY: Representation of Ion -- Protein Interactions Using the Drude Polarizable Force-Field​. ​J Phys Chem B​ 2015, ​119​:9401 -- 9416. 34. Yu H, Whitfield TW, Harder E, Lamoureux G, Vorobyov I, Anisimov VM, MacKerell AD, Roux B: ​Simulating Monovalent and Divalent Ions in Aqueous Solution Using a Drude Polarizable Force Field​. ​J Chem Theory Comput​ 2010, ​6​:774 -- 786. 35. Ponder JW, Wu C, Ren P, Pande VS, Chodera JD, Schnieders MJ, Haque I, Mobley DL, Lambrecht DS, DiStasio RA, et al.: ​Current Status of the AMOEBA Polarizable Force Field​. ​J Phys Chem B​ 2010, ​114​:2549 -- 2564. 36. Jing Z, Liu C, Qi R, Ren P: ​Many-body effect determines the selectivity for Ca ​2+​ and Mg ​2+​ in proteins​. ​Proc Natl Acad Sci​ 2018, ​115​:E7495 -- E7501. 37. Lin F-Y, Lopes PEM, Harder E, Roux B, MacKerell AD: ​Polarizable Force Field for Molecular Ions Based on the Classical Drude Oscillator​. ​J Chem Inf Model​ 2018, 58​:993 -- 1004. and with Nucleic Acid Moieties For a Polarizable Force Field Based on the Classical Drude Oscillator Model​. ​J Phys Chem B​ 2016, ​120​:11436 -- 11448. Drude Oscillator: I. Refinement Using Quantum Mechanical Base Stacking and Conformational Energetics​. ​J Chem Theory Comput​ 2017, ​13​:2053 -- 2071. Drude Oscillator: II. Microsecond Molecular Dynamics Simulations of Duplex DNA​. ​J Chem Theory Comput​ 2017, ​13​:2072 -- 2085. 41. Lemkul JA, MacKerell AD: ​Polarizable force field for RNA based on the classical drude oscillator: Polarizable Force Field for RNA Based on the Classical Drude Oscillator​. ​J Comput Chem​ 2018, ​39​:2624 -- 2646. 42. Zhang C, Lu C, Wang Q, Ponder JW, Ren P: ​Polarizable Multipole-Based Force Field for Dimethyl and Trimethyl Phosphate​. ​J Chem Theory Comput​ 2015, ​11​:5326 -- 5339. 43. Mao Y, Demerdash O, Head-Gordon M, Head-Gordon T: ​Assessing Ion -- Water Interactions in the AMOEBA Force Field Using Energy Decomposition Analysis of Electronic Structure Calculations​. ​J Chem Theory Comput​ 2016, ​12​:5422 -- 5437. 44. Jing Z, Qi R, Liu C, Ren P: ​Study of interactions between metal ions and protein model compounds by energy decomposition analyses and the AMOEBA force field​. J Chem Phys​ 2017, ​147​:161733. 45. Wang Q, Rackers JA, He C, Qi R, Narth C, Lagardere L, Gresh N, Ponder JW, Piquemal J-P, Ren P: ​General Model for Treating Short-Range Electrostatic Penetration in a Molecular Mechanics Force Field​. ​J Chem Theory Comput​ 2015, 11​:2609 -- 2618. 46. Deng S, Wang Q, Ren P: ​Estimating and modeling charge transfer from the SAPT induction energy​. ​J Comput Chem​ 2017, ​38​:2222 -- 2231. 47. Rick SW: ​A polarizable, charge transfer model of water using the drude oscillator: Influence of Charge Transfer on the Drude Oscillator​. ​J Comput Chem​ 2016, 37​:2060 -- 2066. 39. Lemkul JA, MacKerell AD: ​Polarizable Force Field for DNA Based on the Classical 40. Lemkul JA, MacKerell AD: ​Polarizable Force Field for DNA Based on the Classical 22 51. Walters ET, Mohebifar M, Johnson ER, Rowley CN: ​Evaluating the London 54. Hol WGJ, Duijnen PT van, Berendsen HJC: ​The α-helix dipole and the properties of 55. Huang J, MacKerell AD: ​Induction of Peptide Bond Dipoles Drives Cooperative 56. Babu MM, van der Lee R, de Groot NS, Gsponer J: ​Intrinsically disordered proteins: 57. Huang J, MacKerell AD: ​Force field development and simulations of intrinsically 58. Robustelli P, Piana S, Shaw DE: ​Developing a molecular dynamics force field for 48. Awoonor-Williams E, Rowley CN: ​The hydration structure of methylthiolate from QM/MM molecular dynamics​. ​J Chem Phys​ 2018, ​149​:045103. 49. Lin F-Y, MacKerell AD: ​Polarizable Empirical Force Field for Halogen-Containing Compounds Based on the Classical Drude Oscillator​. ​J Chem Theory Comput​ 2018, 14​:1083 -- 1098. 50. Mohebifar M, Johnson ER, Rowley CN: ​Evaluating Force-Field London Dispersion Coefficients Using the Exchange-Hole Dipole Moment Model​. ​J Chem Theory Comput​ 2017, ​13​:6146 -- 6157. Dispersion Coefficients of Protein Force Fields Using the Exchange-Hole Dipole Moment Model​. ​J Phys Chem B​ 2018, ​122​:6690 -- 6701. 52. Cole DJ, Vilseck JZ, Tirado-Rives J, Payne MC, Jorgensen WL: ​Biomolecular Force Field Parameterization via Atoms-in-Molecule Electron Density Partitioning​. ​J Chem Theory Comput​ 2016, ​12​:2312 -- 2323. 53. Visscher KM, Geerke DP: ​Deriving Force-Field Parameters from First Principles Using a Polarizable and Higher Order Dispersion Model​. ​J Chem Theory Comput 2019, ​15​:1875 -- 1883. proteins​. ​Nature​ 1978, ​273​:443 -- 446. Helix Formation in the (AAQAA)3 Peptide​. ​Biophys J​ 2014, ​107​:991 -- 997. regulation and disease​. ​Curr Opin Struct Biol​ 2011, ​21​:432 -- 440. disordered proteins​. ​Curr Opin Struct Biol​ 2018, ​48​:40 -- 48. both folded and disordered protein states​. ​Proc Natl Acad Sci​ 2018, 115​:E4758 -- E4766. MacKerell AD: ​CHARMM36m: an improved force field for folded and intrinsically disordered proteins​. ​Nat Methods​ 2017, ​14​:71 -- 73. simulations​. ​Curr Opin Struct Biol​ 2018, ​49​:129 -- 138. Structure Refinement, Protein Folding, and Intrinsically Disordered Proteins Using Polarizable Force Fields​. ​J Phys Chem Lett​ 2018, ​9​:7110 -- 7116. 62. Ngo VA, Fanning JK, Noskov SY: ​Comparative Analysis of Protein Hydration from MD simulations with Additive and Polarizable Force Fields​. ​Adv Theory Simul​ 2019, 2​:1800106. 63. Hazel AJ, Walters ET, Rowley CN, Gumbart JC: ​Folding free energy landscapes of β-sheets with non-polarizable and polarizable CHARMM force fields​. ​J Chem Phys​ 2018, ​149​:072317. Computational and Experimental Studies of Inhibitor Design for Aldolase A​. ​J Phys Chem B​ 2019, ​123​:6034 -- 6041. with Additive and Polarizable Free Energy Simulations​. ​Biophys J​ 2018, 114​:1091 -- 1102. 65. Panel N, Villa F, Fuentes EJ, Simonson T: ​Accurate PDZ/Peptide Binding Specificity 59. Huang J, Rauscher S, Nawrocki G, Ran T, Feig M, Groot BL de, Grubmüller H, 60. Nerenberg PS, Head-Gordon T: ​New developments in force fields for biomolecular 61. Wang A, Zhang Z, Li G: ​Higher Accuracy Achieved in the Simulations of Protein 64. Qi R, Walker B, Jing Z, Yu M, Stancu G, Edupuganti R, Dalby KN, Ren P: 23 71. Sun R-N, Gong H: ​Simulating the Activation of Voltage Sensing Domain for a 72. Dhakshnamoorthy B, Rohaim A, Rui H, Blachowicz L, Roux B: ​Structural and 73. Kratochvil HT, Carr JK, Matulef K, Annen AW, Li H, Maj M, Ostmeyer J, Serrano AL, 69. Flood E, Boiteux C, Lev B, Vorobyov I, Allen TW: ​Atomistic Simulations of 66. Welborn VV, Head-Gordon T: ​Fluctuations of Electric Fields in the Active Site of 67. Griffiths TM, Oakley AJ, Yu H: ​Atomistic Insights into Photoprotein Formation: the Enzyme Ketosteroid Isomerase​. ​J Am Chem Soc​ 2019, ​141​:12487 -- 12492. Computational Prediction of the Properties of Coelenterazine and Oxygen Binding in Obelin​. ​J Comput Chem​ 2019, doi:10.1002/jcc.26125. 68. Torabifard H, Cisneros GA: ​Computational investigation of O ​2​ diffusion through an intra-molecular tunnel in AlkB; influence of polarization on O ​2​ transport​. ​Chem Sci​ 2017, ​8​:6230 -- 6238. Membrane Ion Channel Conduction, Gating, and Modulation​. ​Chem Rev​ 2019, 119​:7737 -- 7832. 70. Peng X, Zhang Y, Chu H, Li Y, Zhang D, Cao L, Li G: ​Accurate Evaluation of Ion Conductivity of the Gramicidin A Channel Using a Polarizable Force Field without Any Corrections​. ​J Chem Theory Comput​ 2016, ​12​:2973 -- 2982. Voltage-Gated Sodium Channel Using Polarizable Force Field​. ​J Phys Chem Lett 2017, ​8​:901 -- 908. functional characterization of a calcium-activated cation channel from Tsukamurella paurometabola​. ​Nat Commun​ 2016, ​7​:12753. Raghuraman H, Moran SD, et al.: ​Instantaneous ion configurations in the K ​+​ ion channel selectivity filter revealed by 2D IR spectroscopy​. ​Science​ 2016, 353​:1040 -- 1044. 74. Riahi S, Rowley CN: ​Why Can Hydrogen Sulfide Permeate Cell Membranes? ​J Am Chem Soc​ 2014, ​136​:15111 -- 15113. 75. Loco D, Lagardère L, Cisneros GA, Scalmani G, Frisch M, Lipparini F, Mennucci B, Piquemal J-P: ​Towards large scale hybrid QM/MM dynamics of complex systems with advanced point dipole polarizable embeddings​. ​Chem Sci​ 2019, ​10​:7200 -- 7211. 76. Boulanger E, Thiel W: ​Toward QM/MM Simulation of Enzymatic Reactions with the Drude Oscillator Polarizable Force Field​. ​J Chem Theory Comput​ 2014, 10​:1795 -- 1809. 77. König G, Pickard FC, Huang J, Thiel W, MacKerell AD, Brooks BR, York DM: ​A Comparison of QM/MM Simulations with and without the Drude Oscillator Model Based on Hydration Free Energies of Simple Solutes​. ​Molecules​ 2018, ​23​:2695. 78. Ganguly A, Boulanger E, Thiel W: ​Importance of MM Polarization in QM/MM Studies of Enzymatic Reactions: Assessment of the QM/MM Drude Oscillator Model​. ​J Chem Theory Comput​ 2017, ​13​:2954 -- 2961. 79. Inakollu VSS, Yu H: ​A systematic benchmarking of computational vibrational spectroscopy with DFTB3: Normal mode analysis and fast Fourier transform dipole autocorrelation function​. ​J Comput Chem​ 2018, ​39​:2067 -- 2078. 80. Semrouni D, Sharma A, Dognon J-P, Ohanessian G, Clavaguéra C: ​Finite Temperature Infrared Spectra from Polarizable Molecular Dynamics Simulations​. J Chem Theory Comput​ 2014, ​10​:3190 -- 3199. Deprotonated Dicarboxylic Acids: IRMPD Spectroscopy and Empirical Valence-Bond Modeling​. ​ChemPhysChem​ 2019, ​20​:803 -- 814. 81. Thaunay F, Calvo F, Nicol E, Ohanessian G, Clavaguéra C: ​Infrared Spectra of 82. Thaunay F, Dognon J-P, Ohanessian G, Clavaguéra C: ​Vibrational mode assignment 24 of finite temperature infrared spectra using the AMOEBA polarizable force field​. Phys Chem Chem Phys​ 2015, ​17​:25968 -- 25977. 83. Giovannini T, Grazioli L, Ambrosetti M, Cappelli C: ​Calculation of IR Spectra with a Fully Polarizable QM/MM Approach Based on Fluctuating Charges and Fluctuating Dipoles​. ​J Chem Theory Comput​ 2019, ​15​:5495 -- 5507. 25
1906.05043
2
1906
2019-10-25T14:05:47
Mechanics of allostery: contrasting the induced fit and population shift scenarios
[ "physics.bio-ph", "cond-mat.soft" ]
In allosteric proteins, binding a ligand can affect function at a distant location, for example by changing the binding affinity of a substrate at the active site. The induced fit and population shift models, which differ by the assumed number of stable configurations, explain such cooperative binding from a thermodynamic viewpoint. Yet, understanding what mechanical principles constrain these models remains a challenge. Here we provide an empirical study on 34 proteins supporting the idea that allosteric conformational change generally occurs along a soft elastic mode presenting extended regions of high shear. We argue, based on a detailed analysis of how the energy profile along such a mode depends on binding, that in the induced fit scenario there is an optimal stiffness $k_a^*\sim 1/N$ for cooperative binding, where $N$ is the number of residues involved in the allosteric response. We find that the population shift scenario is more robust to mutation affecting stiffness, as binding becomes more and more cooperative with stiffness up to the same characteristic value $k_a^*$, beyond which cooperativity saturates instead of decaying. We confirm numerically these findings in a non-linear mechanical model. Dynamical considerations suggest that a stiffness of order $k_a^*$ is favorable in that scenario as well, supporting that for proper function proteins must evolve a functional elastic mode that is softer as their size increases. In consistency with this view, we find a significant anticorrelation between the stiffness of the allosteric response and protein size in our data set.
physics.bio-ph
physics
Mechanics of allostery: contrasting the induced fit and population shift scenarios Riccardo Ravasio,1 Solange Marie Flatt,1 Le Yan,2, ∗ Stefano Zamuner,1 Carolina Brito,3 and Matthieu Wyart1, † 1Institute of Physics, ´Ecole Polytechnique F´ed´erale de Lausanne, CH-1015 Lausanne, Switzerland 2Kavli Institute for Theoretical Physics, University of California, Santa Barbara, CA 93106, USA 3Instituto de F´ısica, Universidade Federal do Rio Grande do Sul, CP 15051, 91501-970 Porto Alegre RS, Brazil In allosteric proteins, binding a ligand can affect function at a distant location, for example by changing the binding affinity of a substrate at the active site. The induced fit and population shift models, which differ by the assumed number of stable configurations, explain such cooperative binding from a thermodynamic viewpoint. Yet, understanding what mechanical principles constrain these models remains a challenge. Here we provide an empirical study on 34 proteins supporting the idea that allosteric conformational change generally occurs along a soft elastic mode presenting extended regions of high shear. We argue, based on a detailed analysis of how the energy profile along such a mode depends on binding, that in the induced fit scenario there is an optimal stiffness a ∼ 1/N for cooperative binding, where N is the number of residues. We find that the population k∗ shift scenario is more robust to mutation affecting stiffness, as binding becomes more and more cooperative with stiffness up to the same characteristic value k∗ a, beyond which cooperativity satu- rates instead of decaying. We confirm numerically these findings in a non-linear mechanical model. Dynamical considerations suggest that a stiffness of order k∗ a is favorable in that scenario as well, supporting that for proper function proteins must evolve a functional elastic mode that is softer as their size increases. In consistency with this view, we find a fair anticorrelation between the stiffness of the allosteric response and protein size in our data set. Significance statement: Many proteins are al- losteric: binding a ligand affects their activity at a distant site. Understanding the principles allowing for such an action at a distance is both of fundamental and practical importance. From the thermodynamic viewpoint, two models have been proposed, according to which binding creates a new stable configuration or instead shifts the thermal equilibrium between existing states. We perform a mechanical analysis of these models and show that they are not equally robust to mutations. We argue that in both cases function can most properly occur along a soft elastic mode, whose stiffness decreases rapidly with protein size. We show data on 34 proteins substantiating this result, supporting a new principle for allosteric design. Many proteins are allosteric: binding a ligand at one or several allosteric sites can regulate function at a distant site, a long-range communication often accompanied by large conformational changes [1, 2]. There is a consider- able interest in predicting the amino acids involved in this communication, or "allosteric pathway", from structure or sequence data [3, 4], since they can be used as tar- gets for drug design [5]. Yet, understanding the physical principles underlying such action at a distance in pro- teins remains a challenge [6, 7]. From a thermodynamic standpoint, two distinct views have been proposed. In the induced fit scenario, exemplified by the Koshland- N´emethy-Filmer (KNF) model [8], the protein essentially lies in one single state. The latter changes as binding oc- curs, leading to a conformational change. In an energy landscape picture such as that of Fig.1B, it corresponds to a displacement of the energy minimum upon binding. By contrast, in the population shift model, exemplified by the Monod-Wyman-Changeux (MWC) model [9], two states are always present. Their relative stability can change sign upon binding, leading to an average confor- mational change. Although each of these models pre- sumably applies to various proteins, they do not specify which designs allow for efficient action at a distance, and how robust these designs are to mutations [10]. In several proteins -- see below for a systematic study -- it has been observed that the allosteric re- sponse induced by binding occurs predominantly along one or few vibrational modes [11 -- 15]. This result sup- ports that in at least some proteins elasticity -- possi- bly non-linear -- is an appropriate language to describe allostery (in contrast to intrinsically disordered proteins that may be considered more as liquids than solids, for which the analysis proposed here would not hold). Very recently, there has been a considerable effort to use in- silico evolution [16, 17] to study how linear elastic ma- terials can evolve to accomplish an allosteric task [18 -- 26]. In general, binding a ligand locally distorts the pro- tein, which is modelled by imposing local displacements at some site, generating an extended elastic response that in turn determines fitness (chosen specifically to accom- plish a given task). These models fall into the induced fit scenario, since in the framework of linear elasticity there is always a single minimum of energy. A particu- larly key allosteric function within proteins is the amount of cooperative binding, defined as the change of binding energy of a substrate at the active site caused by bind- ing a ligand at the allosteric site. Materials optimized to display such a cooperativity over long-distances develop a single extended "mechanism" -- a soft elastic mode, such as the motion of closing scissors -- connecting the two binding sites [21]. It is found that the stiffness ka 9 1 0 2 t c O 5 2 ] h p - o i b . s c i s y h p [ 2 v 3 4 0 5 0 . 6 0 9 1 : v i X r a (i.e. the curvature of the energy) of this mode cannot be too large nor too small for cooperativity to occur, and a ∼ 1/N where N that optimal design corresponds to k∗ is the number of particles in the system. If proteins are nearly optimal, mutations stiffening that mode should thus diminish cooperativity. Yet, these predictions are restricted to strictly linear elasticity, an approximation that presumably does not hold in the regime where most protein operate -- certainly not in the population shift scenario. In this work, we show that the population shift and the induced fit models are very different from a mechan- ical perspective. In the former case as well, function can be achieved by developing a mechanism or soft elastic mode, but cooperativity, instead of steadily decreasing, saturates to a constant value once the mode stiffness passes some characteristic value, whose scaling with N a ∼ 1/N . We confirm this prediction in a non-linear is k∗ elastic model of allostery. This result implies that co- operativity is more robust towards mutations increasing stiffness in the population shift scenario. Yet, display- ing a stiffness much larger than k∗ a implies a very long transition time between the two states, and is presum- ably prohibited, suggesting the hypothesis that allosteric proteins function with modes presenting a stiffness near the characteristic value k∗ a in both cases. We test this proposition systematically using X-ray crystallographic data of 34 allosteric proteins. We first confirm that one or a few vibrational modes contribute to the allosteric re- sponse, and introduce a new observable establishing that this response presents unusually extended regions of large shear, as found previously for three proteins [27]. Next we confirm that the characteristic mode stiffness tends to decrease with the propagation length as we expect from the predicted scaling of k∗ a. Finally, we suggest sys- tematic mutational studies to further test how mechanics constrains allostery. GEOMETRICAL INTERPRETATION OF MECHANICAL CONSTRAINTS IN INDUCED FIT ALLOSTERY As sketched in Fig.1A, a protein with two binding sites can be unbound (labeled "00"), bound to a single ligand (labeled "01" or "10") or doubly bound (labeled "11"). We define by E00, E10, E01, E11 the energy of the protein in these four situations (corresponding to the minimum energy of their energy landscape), and choose E00 = 0 as the reference energy. Cooperativity is then defined as: ∆∆E = E10 + E01 − E11 (1) To simplify notation below, we assume a symmetry be- tween the states 01 and 10, in particular E10 = E01. Our qualitative conclusions below however remain valid even if this symmetry does not hold. 2 FIG. 1: (A) A protein with two binding sites can be unbound "00", bound to a single ligand "01" (or "10") or doubly bound "11". (B) Induced fit scenario. Elastic energy of the unbound state E00(x)/ka (in black) re-scaled for visibility by the soft mode stiffness ka, singly bound state E01(x)/ka (in blue) and doubly bound state E11(x)/ka (in red) as a function of the imposed motion x along its soft mode. E00, E10 and E11 correspond to the minima of the black, blue and red curves respectively. In this sketch we have assumed that for a motion x0, the protein shape can accommodate perfectly both ligands without deforming, thus the three energy profiles are identical at that point. Three cases are sketched, depending on the magnitude of the characteristic stiffness of the mode ka. Consider a protein displaying cooperativity thanks to the presence of a soft elastic mode. Let us denote by x a variable indicating the motion along that mode (see our numerical model below for a concrete example in the con- text of a shear design), which varies from zero to unity as the protein undergoes its allosteric response. The energy profile E00(x) of the unbound state follows: E00(x) = kadRa2f (x) where dRa2 =(cid:80) (2) i=1...N d (cid:126)Ri2 is the norm of the al- losteric response dRa(cid:105) ≡ {d (cid:126)Ra(i)} and d (cid:126)Ra(i) is the vec- tor displacement of the amino acid i. We expect that if the protein has N residues, dRa2 is of order N a2, where a is the inter-atomic distance. We will confirm em- pirically the dependence of dRa2 with N below. ka is the mode stiffness and f (x) some function of order unity such that f(cid:48)(cid:48)(x = 0) = 1. For a purely linear elastic ma- terial, f (x) = x2/2. More generally for the induced fit scenario, f (x) presents a single minimum, as illustrated in Fig.1B. In [21] it was argued that if linear elasticity applies, then there is an optimal stiffness k∗ a for cooper- ativity. We present a geometrical interpretation of this result, which extends it to the induced fit scenario in general, and will be useful to explain why the popula- tion shift scenario behaves differently. Local chemical interactions which lead to identical binding energies for 10000111ACAB$k$L^{2/3}$IIIIIIIIIIII$k$$k$$L^{2/3}$$L^{2/3}$$\Huge E_{\rm coop}$$\Huge E_{\rm coop}$ the inactive and active states do not affect cooperativ- ity, and thus need not be considered here. For concrete- ness, we assume that after some motion x0 = 1 along that mode, the protein shape can accommodate perfectly both ligands without deforming. It implies that the en- ergy profile of the bound states E10(x) and E11(x) sat- isfy E10(x0) = E11(x0) = E00(x0), as pictured in Fig.1B. However as x departs from x0, the protein shape does not match the ligands, imposing an elastic strain at the binding sites leading to an increase of elastic energy in the protein, that will trigger motion along other elastic modes. Assuming that the ligands are rigid and that each binding site involves on the order n0 atoms that move by a distance a as the protein undergoes its allosteric re- sponse, we have E01(x)− E00(x) = n0ka2g(x− x0) where k characterizes the stiffness of inter-atomic interactions and is large in comparison with the one of soft modes, i.e. k (cid:29) ka. Here g is some dimensionless function vanishing quadratically in zero -- but possibly non-linear at large arguments. E11, E01 can be computed as the minimum of the curves E11(x) and E01(x) respectively, from which the cooperativity ∆∆E is readily computed. Two extreme cases occur, illustrated in Fig.1B: (i) if kaN (cid:28) n0k (Fig.1B.I), as x moves away from x0, the elastic energy induced by binding n0ka2g(x − x0) is very significant in comparison to the mode energy E00(x) ∼ kaN a2f (x). Thus both E10(x) and E11(x) have a sharp minimum near x0, with E11 ≈ E10 ≈ kadRa2f (1). Thus ∆∆E = E10 + E01 − E11 ≈ E10 ≈ kadRa2f (1), which vanishes as ka → 0. (ii) If kaN (cid:29) n0k is very large (Fig.1B.III), n0ka2g(x− x0) is small in comparison to E00(x): E00(x), E11(x) and E01(x) are very close to each other, and must thus all present a minimum near x = 0. Thus binding does not trigger motion along the soft mode, whose presence is useless. No extended modes couples the two binding sites and E11 ≈ E01 + E10, leading to ∆∆E → 0 as ka → ∞. (iii) Optimal cooperativity is thus found at some inter- a ∼ n0k/N , corresponding to Fig.1B.II. Note mediary k∗ that the present argument for an optimal k∗ a does not require the energy profile f (x) to be an exact parabola, as long as it is monotonically growing in both directions around its minimum. Mechanical aspects of the population shift model MWC model We recall some aspects of the MWC model. To sim- plify notation, we consider that the protein displays two symmetric binding sites as illustrated in Fig.1A. The protein is assumed to lie in two possible distinct con- figurations, "Inactive" (In) and "Active" (Ac). In the absence of binding, we take the energy of the inactive 3 FIG. 2: Cooperative energy ∆∆E as function of the energy cost of binding a ligand E0 is sketched for the MWC model. The maximal cooperative energy that the system can reach is ∆∆E∗. 10 = EAc 01 = EAc state as our reference (i.e. EIn 00 = 0) and denote the en- ergy of the active state EAc 00 = E0. We assume that the active configuration has a well-suited geometry to bind each ligand, thus no elastic energy is spent for binding and EAc 11 = E0. By contrast, we assume that in the inactive state binding costs some energy ∆E, leading to EIn 11 = 2∆E. The last assumption of additivity of binding energies within a given configuration, i.e. for a frozen mode amplitude, is expected to be accurate if the two binding sites are distant enough, since no elastic coupling between them is expected in that case. 01 = ∆E and EIn 10 = EIn For each binding situation, the configuration (inactive or active) chosen is the one with the smallest energy, e.g. E01 = min(EIn 01 ). Computing the cooperativity one finds three different cases: 01 , EAc (i) if E0 < ∆E, then the binding of one ligand is suf- ficient to drive the system in the active state, implying E10 = E11 = E0 and ∆∆E = E0; (ii) if ∆E < E0 < 2∆E, then the binding of one ligand is not sufficient and two ligands have to bind in order to drive the system in the active state. Consequently, ∆∆E = 2∆E − E0; (iii) if E0 > 2∆E, then ∆∆E = 0 since the system stays in the inactive state even if two ligands are bound. The sketch of this behavior is shown in Fig.2, illustrating that the maximum cooperativity is found for E0 = ∆E. Mechanical consideration on the MWC model Our observations (see the empirical section below) in- dicate that a significant fraction of allosteric proteins op- erate mainly along one normal mode of the elastic en- ergy, supporting the idea that in these cases a favored path connects the inactive and active configurations. We expand the energy (in the absence of ligand) in terms of the motion x along that path (in this two-states case, it is more convenient to chose a coordinate x varying be- tween −1 and +1 as the protein undergoes its allosteric response). Note that if non-linearities are present, this 4 8 kadRa2. is maximal at b = bc, where one finds E0 = 3 (iv) For small b, E0 (cid:39) 2kadRa2b. In what follows we focus on the case b < bc where the population shift model lies. Next, we assume that the active configuration matches the shape of both ligands, so that binding events in that state cost no energy. However, by moving away from the configuration the shapes of the protein and ligands do not match anymore and the protein needs to deform elastically near the binding site. Again assuming that each binding site involves of order n0 atoms, which move by a distance of order a from the active to the inac- tive state, we have a binding energy E10(x) − E00(x) = n0ka2g(x − xAc) where xAc ≈ 1 is the location of the active state along the path. This energy is exemplified by the difference between the blue and black curves in Fig.3A. Thus the binding energy in the inactive state follows ∆E = n0ka2g(xIn − xAc) ≈ 4n0ka2, which is independent of the protein size and mode stiffness ka. As explained before, if the two binding sites are distant enough, for a given mode amplitude the elastic costs E11(x) − E00(x) = of binding will simply add up: 2n0ka2g(x − xAc) as shown in red in Fig.3A. How the mode stiffness constrains cooperativity To quantify this constraint we define the maximal co- operativity over all possible tilts b given ka: ∆∆E∗ ≡ maxb {∆∆E(ka, b)}. We find two regimes: (i) If N ka (cid:28) n0k, then the elastic costs associated with binding are very large compared to E0. Both E10(x) and E11(x) are peaked close to xAc, as illustrated in Fig. 3B.a. Thus E0 < ∆E, implying ∆∆E = E0 ac- cording to Fig. 2, which is maximized at b = bc leading 8 kadRa2 ∼ kaN a2. Thus ∆∆E vanishes to ∆∆E = 3 linearly at small ka, as illustrated in Fig.3B. This result is qualitatively similar to the induced fit case, for which ∆∆E also vanishes linearly as shown in Fig.1.B. (ii) If N ka (cid:29) n0k, then the elastic cost is very small in comparison to E0: E00(x), E10(x) and E11(x) are almost identical as illustrated in Fig.3B.c. In that regime, coop- erativity is optimized by chosing a small tilt fixing E0 to ∆E according to Fig. 2, implying ∆∆E = ∆E ∼ n0ka2 which is independent of ka. This plateau behaviour is represented in Fig. 3B, and appears at k∗ a ∼ n0k/N . This result represents a fundamental difference with the induced fit case, for which a large stiffness destroys cooperativity. Indeed in the induced fit scenario a large stiffness implies that the minimal energy is always found for x ≈ 0 as illustrated in Fig. 1B.III, implying that bind- ing does not move the protein along that mode, which is thus useless. This state of affairs is ultimately a geo- metric necessity stemming from the fact that the three curves E00(x), E10(x) and E11(x) must be very close to each other in that regime, and each present a single min- A B FIG. 3: (A) Sketch of the energy profiles E00(x) (black), E10(x) (blue) and E11(x) (red) between two states active (Ac) and inactive (In) as a function of the motion x along the path connecting them. The energy difference E0 between the inactive and active unbound state, the energy cost of binding one ligand ∆E and the height of the energy barrier Eb are highlighted. (B) Maximal cooperative energy ∆∆E∗ as function of the stiffness of the allosteric response ka, showing (I) linear growth and (II) a a and ka (cid:29) k∗ plateau. energy profiles for ka (cid:28) k∗ a are show in inset. There are obtained from the non-linear elastic model of allostery described below. a, ka ∼ k∗ path is not along a single linear mode, but it bends in phase space. Our analysis below holds independently of such bending. We keep the minimal number of non-linear terms that allow to display two states (a polynomial of four degrees has five parameters, three of them can be fixed by redefining the reference energy, and changing the definition of x by both a multiplicative and additive constants): (cid:18) 1 (cid:19) , (3) E00(x) = kadRa2 x4 − 1 4 x2 + xb 8 where b is a parameter reflecting how the energy profile is tilted towards the inactive state, ka characterizes the stiffness of the mode and dRa2 is the square norm of the allosteric response. A typical profile following Eq.3 is shown in Fig.3A. We denote by inactive the lowest of the two minima and the other one by active. Note that (i) in the case b = 0, Eq.3 describes two identical minima at x = ±1 of stiffness ka, separated by an energy barrier Eb = kadRa2/8. (ii) When the parameter b is positive, we have EAc > EIn 3) where the active state becomes up to b = bc = 1/(3 unstable and only a single stable state is left. (iii) At fixed ka, the energy difference between the two states E0 √ II*kaStiffness kaabc 5 As an illustration, we consider the shear design in which the protein presents two three-dimensional rigid regions connected by a soft planar layer that can easily deform. The rigid regions consist of harmonic springs of stiffness k, shown in blue in Fig.4A. The soft layer consists of anharmonic springs (shown in red in Fig.4A), whose energy vs. extension curve is non-linear, chosen to have the same form as Eq.3. These non-linear springs have a characteristic stiffness kw, and present two stable extensions at which they exert no force, whose relative energy is controlled by some bias bw. These stable ex- tensions are chosen such that two states of the protein as a whole, the inactive and active states shown in Fig.4A, present no contact forces by construction, and are thus local minima of the energy. Finally the protein presents two binding sites, at its top and bottom. At each site, binding imposes that the distance between two nodes (in- dicated by black lines in Fig.4A) equals the distance it naturally presents in the active state, thus favoring it. The details about the construction of the microscopic model are discussed in Supplementary Information. We can now numerically compute the energy profiles E00(x), E01(x) and E11(x) as a function of the motion along the shear mode x, by imposing a shear displace- ment (i.e. a value of x) and letting the entire elastic energy of the material relax, except for that mode. The insets of Fig. 4 show our results. The value of the mode stiffness ka can be extracted from fitting Eq.3 to E00(x) and measuring the displacement norm dRa, and can be increased by increasing kw. From E00(x) one readily computes the energy difference E0 between the inactive and active states, which can be increased by monitoring the microscopic bias bw. From the minima of E00(x), E01(x) and E11(x) one immediately extracts the binding costs ∆E and ∆∆E. Fig.4B shows ∆∆E for two values of ka, as the energy difference E0 is increased. For large ka, ∆∆E passes through a maximum ∆∆E∗ = ∆E, whereas for small ka, ∆∆E is smaller, and its maximum is fixed by the maximal achievable energy difference E0. ∆∆E∗ as a function of ka is shown for different system sizes N in Fig.4C, confirming the presence of two regimes with a a ∼ 1/N as shown in Fig.4D. Overall, cross-over at some k∗ these observations validate our theoretical predictions on the dependence of k∗ a with the number of residues, and on how the stiffness ka qualitatively affects cooperativity. Empirical study of 34 allosteric proteins Allosteric Response We identify a set of 34 allosteric proteins in the Pro- tein Data Bank (PDB), for which both the active (lig- and bound at the allosteric site) and inactive (no lig- and bound at the allosteric site) crystalline X-ray struc- FIG. 4: (A) Sliced view of our three-dimensional mechanical model for population shift allostery. In blue: rigid elastic regions made of harmonic springs of stiffness k, separated by a weak non-linear region made of nonlinear springs (in red) of stiffness kw with kw (cid:28) k. The inactive state (In) favors short springs, while the active state (Ac) long springs, as exemplified by a solid orange spring in the weak band. The location of the binding sites are represented by a solid black spring. (B) ∆∆E as function of E0 for different values of ka for a linear length L = 20, where L3 = N . (C) Maximal cooperative energy ∆∆E∗ as function of ka for L = 6, 10, 20 and L = 30. The kink in these curves define the cross-over stiffness k∗ model N , supporting k∗ a v.s. number of nodes in the a. (D) k∗ a ∼ N−1. imum. Consequently the positions of these minima must be very similar in the three cases, leading to x ≈ 0 in- dependently of binding ligands or not. This geometric necessity vanishes as soon as two minima are present. Note that although ∆∆E asymptotes to a constant for ka (cid:29) k∗ a, the barrier Eb between the inactive and active states grows linearly with ka in that limit. Large barriers would lead to undesirably slow transition rate between states, thus we expect that in practice ka lies reasonably close to k∗ a. A mechanical model for population shift allostery We seek to model that (i) the allosteric response often takes place mainly along a single vibrational mode. (ii) Various architectures can lead to allostery, including the well-known shear [27, 30] or hinge [13] designs and others not falling in these categories [31, 32]. Such a diversity is also found in in silico evolution schemes [21]. Yet, such synthetic architectures always present soft extended re- gions where most of the strain (i.e. relative motion) is located. Such an observation was made in a few proteins [27] and will be generalized below. For proteins present- ing two stable configurations, we expect these regions to present two possible ways of locally stacking amino acids well. InAcRTAcAc00011011ΕΕE*~ kaΕ∗ACDB tures are available. Their PDB identifiers are taken from [27, 33] and reported in Supplementary Information. The set is diverse in functionality, and includes enzymes (13), G-proteins (10), Kinases (3), response regulators (3), DNA-binding proteins (4) and the Human Serum Albu- min, among which 12 protein complexes are present. We can thus estimate the allosteric response dRa(cid:105) as the displacement field between the inactive and active struc- tures (after having aligned them via the software Py- mol 2.1.1 [34]). Here we focus on the motion of the N amino-acids, located by the position of their α-carbon. As an illustration, the allosteric response of a given pro- tein (the elongation factor Tu) is shown in black arrows in Fig. 5A.v From the allosteric response dRa(cid:105), one can readily estimate: (i) the magnitude of the displacement dRa2. (ii) The fraction of the protein involved in the response. For any displacement field, this fraction is usu- ally estimated via the participation ratio [36]: N(cid:80)N dRa2 i=1 dRa(i)4 . P = (4) (iii) A measure of how much relative displacement takes place around atom i. Following [27] we consider the shear pseudo-energy Esh(i) quantifying the amount of strain -- essentially a measure of the relative displacement between adjacent atoms -- at residue i, whose precise definition is given in Supplementary Information. Esh(i) = 0 indi- cates that the protein moves as a rigid body near atom i, and by contrast is large where atoms slide rapidly past each other. Esh(i) is shown in color in Fig. 5A for the protein Tu, illustrating that two parts of the protein are rigidly moving (and counter-rotate), while the central re- gion displays significant pseudo-energy Esh(i), which is reminiscent of a hinge design. Elastic Networks Analysis To estimate protein elasticity we use elastic network models (ENM) [37], in which harmonic springs of iden- tical stiffness are placed between all N α-carbons laying below a chosen cutoff radius Rc ∈ [8 − 12] A. ENM is obviously a crude approximation of real atomic interac- tions. Yet, its simplicity allows for the systematic study of various proteins, and it has been successful in cap- turing normal modes relevant for the function of some proteins. The dependence of the results on the value of Rc are indicated by error bars in Fig. 5B and discussed in Supplementary Information for Fig. 6B. This proce- dure defines an elastic energy from which the matrix of the second derivatives, i.e. the Hessian matrix H, can be computed. From H, one can readily estimate the stiff- ness ka of the allosteric response as the curvature of the elastic energy in that direction: ka = (cid:104)dRaHdRa(cid:105) dRa2 , 6 (5) Finally, the eigenvectors of H define the 3N vibrational modes of the protein {vi(cid:105)}i=1...3N . Following [11 -- 15], the overalp qi between the allosteric response and mode i characterizes their similarity (qi = 1 implies that there are identical): (cid:112)(cid:104)dRadRa(cid:105)(cid:104)vivi(cid:105) . (cid:104)dRavi(cid:105) qi = (6) Geometry of the allosteric response Fig. 5B reports the maximal overlap q∗ ≡ maxi qi, as a function of the participation ratio P . Our observation in- dicate that q∗ is in general large (in half of the case larger than 0.45), supporting further that allostery indeed oc- curs mainly along one mode [11 -- 15]. Interestingly, this effect is stronger when most sites of the protein are in- volved in the allosteric response (P large). Interestingly, some protein complexes (for example ATCase, see the values of q∗ in Table 1 of Supplementary Material) also display a large projection of their allosteric response on a single mode, supporting that mechanical considerations can be valuable in such cases as well. We now provide systematic data supporting that the allosteric response presents extended regions of large shear energy [27]. More specifically, we argue that while some vibrational modes can present significant shear (e.g. localized modes capturing the motion of dandling loops) and other can be extended (such as plane-wave-like modes), the allosteric response is unique in presenting both aspects, thus revealing a specific design principle. To quantify this effect, we introduce the quantity, that can be defined on any displacement field: i Esh(i)2)1/2. log Γ ≡ [γ log10(P) + log10(Esh)] Esh = ((cid:80) (7) where Esh is the total magnitude of the shear energy, i.e. log Γ is large if the dis- placement is extended and if the shear energy is large. The factor γ characterizes the trade-off between these two features. Here we choose γ = 3.5 reflecting the fact that for vibrational modes, we find that P varies about 3.5 times less in relative terms than Esh as shown in Supplementary Information. Thus for γ = 3.5, the spatial extension and the amount of shear equally affect log Γ. Fig. 5C shows log Γ averaged over the 17 proteins with q∗ > 45% for the allosteric response (yellow line), the mode with maximum overlap (blue line) and the first 75 vibrational modes (having subtracted the one with largest overlap) as function of the mode rank. We find 7 A B C FIG. 5: (A) Allosteric response (black arrows) of elongation factor Tu corresponding to the displacement between the inactive state (where GDP is bound at the allosteric site) and active state (where GTP is bound at the allosteric site and the aminoacyl-tRNA can bind at the active site). The phosphate binding loop (allosteric site) is highlighted in green, while the active site is at the interface between the GDP binding domain and the two other domains, one residue of which is highlighted in blue [35]. The shear is encoded in both the color and the thickness of the structure in a logarithmic scale, red corresponds to large shear. The allosteric response is similar to that of a hinge. (B) Maximal overlap q∗ as function of the participation ratio P . (C) The observable log Γ quantifying how modes are both extended and present large shear energy (see main text for definition) averaged over the proteins with overlap larger than 45%, as function of the mode rank for (i) the allosteric response , (ii) the modes with largest overlap and (iii) the first 75 modes (having subtracted the one with largest overlap) as indicated in legend. The green circles correspond to the rank of the mode with largest overlap. The shaded region highlights the range where these modes fall. that Γ is typically 160 times larger for the allosteric re- sponse than for vibrational modes, a very significant dif- ference underlying the specific geometry of the allosteric response. of ENM to build the Hessian matrix. CONCLUSION Scaling of response stiffness ka with protein size We can now test our conjecture that the allosteric stiff- a ∼ 1/N where cooperativity satu- ness ka is close to k∗ rates, which implies in particular an anti-correlation be- tween ka and protein size. In our theoretical estimate of k∗ a, we have assumed that the allosteric response mag- dRa2 ∼ N . nitude is linear in the protein size, i.e. It is a natural assumption since the larger the protein, the more likely its response involves many residues. The relationship between these two quantities is tested in 34 proteins in Fig.6A. We indeed find a strong correlation between the logarithms of dRa and N (Pearson coef- ficient r = 0.76 with p-value p = 1.64 × 10−7. Pearson coefficients are computed on the logarithmic values via the Matlab R2017b function corr ). Overall, data are consistent with our assumption of proportionality. Finally, we plot the allosteric response stiffness ka mea- sured according to Eq.S3 in terms of N for all proteins in Fig. 6B. A key finding is the fair anti-correlation be- tween the logarithms of these two quantities (Pearson coefficient r = −0.64 with p-value p = 4.00× 10−5), sup- porting the idea that larger allosteric proteins need to evolve a softer elastic mode to accomplish function, as expected from our analysis. The signal that we observe is encouraging, especially given the possible sources of noise in the analysis, among which, in particular, the use We have provided systematic evidence that the al- losteric response occurs along one soft elastic mode, and we have introduced a novel observable Γ to establish that this response generally displays unusually extended re- gions of high shear strain. These observations support that for many proteins, elasticity is a useful starting point to describe allostery. We have revisited the two classical thermodynamic models of allostery from this perspective, and provided a detailed study of how the energy profile along the soft mode evolves with binding. We find that induced fit and population shift models qualitatively dif- fer. In the induced fit model, there is an optimal stiff- a ∼ 1/N associated to that mode beyond which ness k∗ the cooperative binding energy eventually decreases to zero. The population shift model is more robust to mu- a ∼ 1/N simply marks a tations affecting stiffness, and k∗ cross-over beyond which cooperativity saturates and the transition time between configurations rapidly explodes. We introduced a novel non-linear elastic model for al- lostery supporting these views. Our key result is that proper function is achieved if proteins evolve an elastic mode whose softness must rapidly decrease with size, a prediction supported by the anti-correlation observed be- tween these quantities. Systematic mutation scan on one single protein, in which binding assays to measure cooperativity are com- bined with single molecule experiments or ultrafast laser pulses to estimate the stiffness of the allosteric response, 0.10.20.30.40.50.600.20.40.60.81020406080-9-8-7-6-5-4 8 A B FIG. 6: (A) Logarithm of the square of the norm of the allosteric response, dRa2, is shown as function of the logarithm of the number of residues, N . The solid line corresponds to dR2 number of residues N . The solid line represents the theoretical prediction k∗ a ∼ N . (B) The logarithm of the stiffness ka as function of the logarithm of the a ∼ 1/N . In both plots r indicates the Pearson coefficient. would be extremely useful to test the predicted relation- ship between these quantities. Molecular dynamics ex- periments could further test how the energy profile along the soft elastic mode evolves with binding. Elucidating such an interplay between thermodynamics and mechan- ics in proteins would be valuable in a variety of tasks, including de novo protein design and the discovery of novel allosteric pathways. We thank P. Barth, B. Bravi, P. De Los Rios, D. Malin- verni, R. B. Phillips and M. Popovi´c for discussions. R.R. is supported by the Swiss National Science Foundation under Grant No. 200021-165509/1. M.W. thanks the Swiss National Science Foundation (Grant No. 200021- 165509) and the Simons Foundation (Grant #454953 Matthieu Wyart). This material is based upon work performed using computational resources supported by the "Center for Scientific Computing at UCSB" and NSF Grant CNS-0960316, by the "High Performance Comput- ing at NY" and by the "High Performance Computing at EPFL". propensities," Nature Communications 7 (2016). [5] Ruth Nussinov and Chung-Jung Tsai, "Allostery in dis- ease and in drug discovery," Cell 153, 293 -- 305 (2013). [6] Shoshana J Wodak, Emanuele Paci, Nikolay V Dokholyan, Igor N Berezovsky, Amnon Horovitz, Jing Li, Vincent J Hilser, Ivet Bahar, John Karanicolas, Ger- hard Stock, et al., "Allostery in its many disguises: From theory to applications," Structure (2019). [7] D Thirumalai, Changbong Hyeon, Pavel I Zhuravlev, and George H Lorimer, "Symmetry, rigidity, and al- losteric signaling: From monomeric proteins to molecular machines," arXiv preprint arXiv:1812.04969 (2018). [8] D. E. Jr. Koshland, G. N´emethy, and D. Filmer, "Com- parison of experimental binding data and theoretical models in proteins containing subunits," Biochemistry 5, 365 -- 385 (1966). [9] Jacque Monod, Jeffries Wyman, and Jean-Pierre Changeux, "On the nature of allosteric transitions: a plausible model," Journal of molecular biology 12, 88 -- 118 (1965). [10] J.J. Hopfield, between "Relation co- operativity and spectra in a model of hemoglobin action," Journal of Molecular Biology 77, 207 -- 222 (1973). structure, ∗ Current address: Chan Zuckerberg Biohub, 499 Illinois † Corresponding St, San Francisco, CA 94158 [email protected], author: [email protected] [1] M. F. Perutz, M. G. Rossmann, Ann F. Cullis, H Muir- head, Georg Will, and A. C. T. North, "Structure of haemoglobin: A three-dimensional fourier synthesis at 5.5-a. resolution, obtained by x-ray analysis," Nature 185, 416 -- 422 (1960). [2] M. F. Perutz, "Stereochemistry of cooperative effects in haemoglobin: Haem -- haem interaction and the problem of allostery," Nature 228, 726 -- 734 (1970). [3] Najeeb Halabi, Olivier Rivoire, Stanislas Leibler, and Rama Ranganathan, "Protein sectors: evolutionary units of three-dimensional structure," Cell 138, 774 -- 786 (2009). [4] Benjamin RC Amor, Michael T Schaub, Sophia N Yali- raki, and Mauricio Barahona, "Prediction of allosteric sites and mediating interactions through bond-to-bond [11] Akio Kitao and Nobuhiro Go, "Investigating protein dy- namics in collective coordinate space," Current Opinion in Structural Biology 9, 164 -- 169 (1999). [12] Ivet Bahar, Burak Erman, Robert L. Jernigan, Ali Rana Atilgan, and David G. Covell, "Collective motions in hiv- 1 reverse transcriptase: Examination of flexibility and en- zyme function," Journal of Molecular Biology 285, 1023 -- 1037 (1999). [13] Chunyan Xu, Dror Tobi, and I Bahar, "Allosteric changes in protein structure computed by a simple me- chanical model: hemoglobin t-r2 transition," Journal of molecular biology 333, 153 -- 168 (2003). [14] Paolo De Los Rios, Fabio Cecconi, Anna Pretre, Giovanni Dietler, Olivier Michielin, Francesco Piazza, and Brice Juanico, "Functional dynamics of PDZ binding domains: a normal-mode analysis," Biophysical journal 89, 14 -- 21 (2005). [15] Wenjun Zheng, Bernard R Brooks, and D Thirumalai, "Low-frequency normal modes that describe allosteric transitions in biological nanomachines are robust to se- quence variations," Proceedings of the National Academy of Sciences 103, 7664 -- 7669 (2006). 22.533.522.533.544.5522.533.5-0.6-0.4-0.200.20.40.6 9 [16] Mathieu Hemery and Olivier Rivoire, "Evolution of spar- sity and modularity in a model of protein allostery," Physical Review E 91, 042704 (2015). [17] Tsvi Tlusty, Albert Libchaber, and Jean-Pierre Eck- mann, "Physical model of the genotype-to-phenotype map of proteins," Phys. Rev. X 7, 021037 (2017). [18] Le Yan, Riccardo Ravasio, Carolina Brito, and Matthieu Wyart, "Architecture and coevolution of allosteric mate- rials," Proceedings of the National Academy of Sciences 114, 2526 -- 2531 (2017). [19] Jason W Rocks, Nidhi Pashine, Irmgard Bischofberger, Carl P Goodrich, Andrea J Liu, and Sidney R Nagel, "Designing allostery-inspired response in mechanical net- works," Proceedings of the National Academy of Sciences 114, 2520 -- 2525 (2017). [20] Holger Flechsig, "Design of elastic networks with evolu- tionary optimized long-range communication as mechan- ical models of allosteric proteins," Biophysical Journal 113, 558 -- 571 (2017). [33] Michael D Daily and Jeffrey J Gray, "Local motions in a benchmark of allosteric proteins," Proteins: Structure, function, and bioinformatics 67, 385 -- 399 (2007). [34] Schrodinger, LLC, "The PyMOL molecular graphics sys- tem, version 1.8," (2015). [35] Morten Kjeldgaard, Poul Nissen, Soren Thirup, and Jens Nyborg, "The crystal structure of elongation factor ef-tu Thermus aquaticus in the GTP conformation," Structure 1, 35 -- 50 (1993). [36] R. J. Bell and P Dean, "Atomic vibrations in vitreous silica," Discuss. Faraday Soc. 50, 55 -- 61 (1970). [37] AR Atilgan, SR Durell, RL Jernigan, MC Demirel, O Ke- skin, and I Bahar, "Anisotropy of fluctuation dynamics of proteins with an elastic network model," Biophysical journal 80, 505 -- 515 (2001). [38] P M Gullett, M F Horstemeyer, M I Baskes, and Fang H, "A deformtion gradient tensor and strain tensors for atomistic simulations," Modelling and Simulation in Ma- terials Science and Engineering 16 (2007). [21] Le Yan, Riccardo Ravasio, Carolina Brito, and Matthieu Wyart, "Principles for optimal cooperativity in allosteric materials," Biophysical Journal 114, 2787 -- 2798 (2018). [22] Sandipan Dutta, Jean-Pierre Eckmann, Albert Libch- and Tsvi Tlusty, "Green function of cor- in a minimal mechanical model of the National (2018), aber, related genes protein evolution," Proceedings Academy http://www.pnas.org/content/115/20/E4559.full.pdf. of Sciences 115, E4559 -- E4568 of [23] Shou-Wen Wang, Anne-Florence Bitbol, and Ned S Wingreen, "Revealing evolutionary constraints on pro- teins through sequence arXiv preprint arXiv:1808.07149 (2018). analysis," [24] Holger Flechsig and Yuichi Togashi, "Designed elastic networks: Models of complex protein machinery," Inter- national journal of molecular sciences 19, 3152 (2018). [25] Barbara Bravi, Riccardo Ravasio, Carolina Brito, and Matthieu Wyart, "Direct coupling analysis of epistasis in allosteric materials," arXiv preprint arXiv:1811.10480 (2018). [26] Jason W Rocks, Henrik Ronellenfitsch, Andrea J Liu, Sidney R Nagel, and Eleni Katifori, "Limits of multi- functionality in tunable networks," Proceedings of the National Academy of Sciences 116, 2506 -- 2511 (2019). [27] Michael R Mitchell, Tsvi Tlusty, and Stanislas Leibler, "Strain analysis of protein structures and low dimension- ality of mechanical allosteric couplings," Proceedings of the National Academy of Sciences , 201609462 (2016). [28] Local chemical interactions which lead to identical bind- ing energies for the inactive and active states do not affect cooperativity, and thus need not be considered here. [29] A polynomial of four degrees has five parameters, three of them can be fixed by redefining the reference energy, and changing the definition of x by both a multiplicative and additive constants. [30] Mark Gerstein, Arthur M Lesk, and Cyrus Chothia, "Structural mechanisms for domain movements in pro- teins," Biochemistry 33, 6739 -- 6749 (1994). [31] Nina M Goodey and Stephen J Benkovic, "Allosteric reg- ulation and catalysis emerge via a common route," Nat Chem Biol 4, 474 -- 482 (2008). [32] Richard N McLaughlin Jr, Frank J Poelwijk, Arjun Ra- man, Walraj S Gosal, and Rama Ranganathan, "The spatial architecture of protein function and adaptation," Nature 491, 138 -- 142 (2012). SUPPLEMENTAL MATERIAL Computing the local strain tensor in a protein 10 In a continuous medium, a motion maps a point (cid:126)X in the reference configuration to a new point (cid:126)x in the current configuration, the strain tensor of the motion can thus be computed as (cid:18) ∂(cid:126)x ∂Xa (cid:19) ab( (cid:126)X) = 1 2 · ∂(cid:126)x ∂Xb − δab , (S4) where a, b labels the spatial dimension. In a discrete medium like proteins which are a collec- tion of atoms (or residues as we consider), computing ↔ Λ = ∂(cid:126)x/∂ (cid:126)X at residue i is not Ideally, for any neighboring residue j the partial derivative straightforward. close enough in space ↔ Λi · ∆ (cid:126)Xij, ∆(cid:126)xij = (S5) where ∆ (cid:126)Xij = (cid:126)Ri0 − (cid:126)Rj0 and ∆(cid:126)xij = (cid:126)Ri − (cid:126)Rj in our setting, where (cid:126)Ri is the position of residue i taken from the X-ray structure. We have nb number of such equa- ↔ tions for Λi are usually over-determined when we consider all neighbors below a certain cutoff distance Rc (we choose Rc1 = 8.5 A for first nearest neighbors and Rc2 = 10.5 A for sec- ond nearest neighbors). Instead of solving Eq. (S5), we define a mean squared error function [38] ↔ Λi when nb neighbors are considered. So M SE(i) = Λi · ∆ (cid:126)Xij)2wj(i), (S6) (cid:88) (∆(cid:126)xij − ↔ j where we have kept a weight function wj(i) of node j contribution to i in general. Specifically, we set as in [27] wj(i) = 1 for all nearest neighbors to i (Rij < Rc1), wj(i) = 1 − Rij − Rc1 Rc2 − Rc1 for Rc1 < Rij < Rc2 and wj = 0 otherwise. By minimizing the mean squared error with respect to ↔ Λi, we have −1 , (cid:88) (cid:18)↔ j ∆(cid:126)xij∆ (cid:126)Xijwj(i) · ∆ (cid:126)Xij∆ (cid:126)Xijwj(i) j ↔  (i) = 1 2 (cid:19) , (S7) (S8) t Λi − ↔ i · ↔ δ Λ A. MECHANICS OF ALLOSTERIC PROTEINS FROM X-RAY STRUCTURAL DATA Dataset We consider 34 allosteric proteins with both inactive and active X-ray structures available [27, 33]. The Pro- tein Data Bank (PDB) entries of these proteins are listed in Table I; the number of common residues between the inactive and active structures and the number of chains that we consider are also reported in the table. We re- move the first two and last two residues for every struc- ture as they correspond to the fluctuating starting and ending points of the protein chain. Elastic network model From the X-ray structures we compute the Hessian ma- trix H using the anisotropic network model introduced in [37]. We consider the positions of residues by looking only at alpha-carbon Cα atoms; Ri for alpha-carbon i. The model assumes a harmonic interaction between two residues at distance smaller than a cutoff distance Rc, Vij = (lij − l0 ij)2 , k 2 (S1) where k is the spring constant (fixed to unity), lij = Ri − Rj is the distance between residues i and j, and l0 ij < Rc is the distance at equilibrium. Building the Hes- sian is then straightforward by taking second derivatives of the potential Vij with respect to the coordinates of residues and evaluated at equilibrium lij = l0 ij, Hiµ,jν = ∂2Vkl ∂Riµ∂Rjν kl = k 2 ∂lkl ∂Riµ ∂lkl ∂Rjν , (S2) where µ, ν = 1, 2, 3 label the spatial dimension of the atoms. The definition of a cutoff distance is an empirical fitting parameter. In practice, we computed Hessian matrices at nine cutoff values equi-distributed in the range [8 − 12] A. We define the stiffness of the allosteric response dRa(cid:105) as (cid:12)(cid:12)(cid:12)(cid:12)lkl=l0 kl (cid:88) kl (cid:88) (cid:88) ↔ Λi = and k∗ a = (cid:104)dRaHdRa(cid:105) (cid:104)dRadRa(cid:105) . (S3) ↔ δ is the identity tensor. where This quantity changes systematically with Rc as shown in Fig.S1A: by increasing Rc, k∗ a also increases, which leads to all points in Fig. S1B moving together in the same direction when changing Rc. In the main text we reported the mean value of k∗ a over the nine choices of Rc, and in Fig. S1B we show the range of variability of a when Rc ∈ [8 − 12] A. k∗ The shear pseudo-energy [27], a vector field whose com- ponents contain a measure of the relative motion of each ↔ residue, can be defined from the strain tensor  (i) com- puted above 3(cid:88) l,m=1 Esh(i) = 1 2 [γlm(i)]2 , ↔ where the local shear tensor δ depends only on the displacement between the two con- ↔ ↔  (i)−(1/3)Tr[  (i)] ↔ γ (i) = formations via the strain tensor ↔  (i) and ↔ δ . Empirical definition of Γ In the main text we introduced the observable Γ com- bining information on the participation ratio (P) and the amount of shear pseudo-energy (Esh) in a given mode as log Γ ≡ [γ log10(P) + log10(Esh)] (S9) with γ = 3.5. In Fig. S2.A we show the range of values of participation ratio (dashed line) compared with the one of shear pseudo-energy (continuous line). The measured relative variation from the two curves is γ∗ (cid:39) 3.46 and we choose γ = 3.5 in the definition of Γ. Fig. S2.B is an alternative way of representing the result discussed in the main text. We consider a scatter plot of log Γ and participation ratio for the three quantities considered (modes with maximum overlap, other modes and allosteric response) and we see that the allosteric response although displaying a large participation ratio it has also a large Γ, meaning that it involves also a considerable amount of shear. B. MICROSCOPIC MODEL To construct a microscopic model with two global min- ima that correspond to the "Inactive" (In) and "Active" (Ac) states, we need to impose that all nonlinear springs to stay exactly at one of their minimum when the net- work is in either In or Ac state. This can be achieved by having each nonlinear spring α following a quartic potential, Eα w = pα 0 + pα 1 xα + pα 2 x2 α + pα 3 x3 α + pα 4 x4 α, (S10) where xα = lα/a, with lα the distance of two particles connected by the interaction α and a a parameter cap- turing the typical local deformation of adjacent particles between In and Ac states. The above compromise can thus be satisfied by choosing the right prefactors pα 0 to pα 4 Given a nonlinear spring in the In and Ac states with rest length, respectively, xIn and xAc, the springs are in a local minimum, if (i) 11 pα 1 + 2pα 2 xAc + 3pα 3 x2 Ac + 4pα 4 x3 Ac = 0, (S11) and (ii) same for xIn. Only the relative value of energy is important, (iii) we can set all pα 0 to zero. (iv) We assume the stiffnesses of the nonlinear springs to be the same, pα 4 = (S12) where kw is the parameter capturing the stiffness of non- linear interactions that has a unit of spring stiffness. (v) Finally, we define kwa2, 1 2 1 (xIn − xAc) + pα pα + pα In − x2 Ac) = kwa2bw, In − x4 2 (x2 4 (x4 Ac) + pα 3 (x3 In − x3 Ac) (S13) where bw is a parameter capturing the energy difference between the Ac and In states. We now have five equa- tions (i-v) for the five parameters (pα 4 ) for each of the nonlinear spring α. Hence, we can define the quartic potentials for all nonlinear springs. 0 to pα Specifically, we consider an elastic network embedded in a Face-Centered-Cubic (FCC) lattice in three dimen- sions (3D) as shown in Fig. S3-A,B. The In and Ac states are connected by a shear mode, which is enabled by ro- tations of two rigid blocks on the left and right against each other, as shown in Fig. S3-C. Such a model can be realized by having harmonic springs with spring stiff- ness k in the two rigid blocks and the nonlinear springs kw (cid:28) k at the boundary of the two. The elastic energy of a linear (strong) spring α is then Eα s = k(lα − l0)2 , 1 2 (S14) where l0 is the rest length of particle separation on the lattice. While, the energy of a nonlinear (weak) spring Eα(cid:48) w obeys Eq. (S10) with parameters solved by the par- ticle distances in two states. The total energy of the system is then obtained by summing the energies over all two kinds of springs, (cid:88) α (cid:88) α(cid:48) E = Eα s + Eα(cid:48) w . (S15) TABLE I: Table with the X-ray structures (PDB identifier) used in the analysis. The number of the common residues and of the considered chains are also reported, along with the participation ratio of the allosteric response and the maximal overlap found between the allosteric response and the normal modes. The PDB identifiers are taken from [33] and [27] (asterisk). 12 arf6 cdc42 rab11 rac1 ras rheb rhoA sec4 IGF-1R Protein type 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. met repressor 11. tet repressor 12. glcN-6-P deaminase 13. EF-Tu 14. Gtα 15. ERK2 16. IRK 17. lac repressor 18. PurR 19. anthranilate synthase 20. chorismate mutase 21. FBPase-1 22. phosphofructokinase 23. PTB1B 24. ATCase∗ 25. hemoglobin 26. NAD-malic enzyme 27. phosphoglycerate DH 28. human serum albinum∗ 29. fixJ 30. DAHP synthase 31. SpoIIAA 32. CheY 33. glycogen phosphorylase 34. ATCase Inactive (PDB) Active (PDB) Common residues 160 183 162 175 164 165 173 152 283 204 190 262 393 310 347 296 536 271 1402 241 323 315 287 A AB A A A A A A A AB A A A A A A AB A ABCD A A A A 2732 ABCDEFGHIJKL ACBD 570 ABCD 2144 ABCD 1580 A 574 AB 238 ABCD 1340 A 106 A 124 1626 AB 2774 ABCDEFGHIJKL Chains Overlap Participation ratio 0.16 0.09 0.07 0.07 0.08 0.1 0.08 0.08 0.03 0.08 0.32 0.26 0.30 0.07 0.06 0.05 0.48 0.34 0.45 0.25 0.04 0.05 0.02 0.55 0.52 0.57 0.53 0.09 0.65 0.55 0.22 0.18 0.24 0.55 0.27 0.33 0.32 0.27 0.44 0.35 0.24 0.23 0.23 0.24 0.47 0.45 0.61 0.36 0.19 0.31 0.71 0.71 0.61 0.54 0.46 0.32 0.48 0.64 0.66 0.72 0.74 0.36 0.69 0.64 0.40 0.34 0.53 0.65 1E0S 1AN0 1OIV 1HH4 4Q21 1XTQ 1FTN 1G16 1P40 1CMB 2TRT 1CD5 1TUI 1TAG 1ERK 1IRK 1TLF 1DBQ 1I7S 2CSM 1EYJ 6PFK 1T48 6AT1 4HHB 1QR6 1PSD 1E78 1DBW 1KFL 1H4Y 3CHY 1GPB 1RAC 2J5X 1NF3 1OIW 1MH1 6Q21 1XTS 1A2B 1G17 1K3A 1CMA 1QPI 1HOT 1EFT 1TND 2ERK 1IR3 1EFA 1WET 1I7Q 1CSM 1EYI 4PFK 1PTY 8AT1 1HHO 1PJ2 1YBA 2BXB 1D5W 1N8F 1H4X 1FQW 7GPB 1D09 A B FIG. S1: (A) We plot the stiffness of the allosteric response ka∗ as function of Rc, where we choose nine equidistributed values of Rc in the range [8 − 12] A. We show the mean of k∗ deviation. We also show the values of k∗ of k∗ response for the data set of allosteric proteins is shown as function of the number of residues in each couple, N . The vertical bars represent the standard deviation associated to the distribution of k∗ a. Increasing the cutoff distance systematically increases the stiffness of all the points. (B) The stiffness associated to the allosteric a versus Rc for the protein with largest (upward triangle) and smallest (downward triangle) values a over all 34 proteins for each Rc (blue circles) with vertical bars given by the standard a resulted from the different values of Rc considered for each protein. 89101112012345610210310-1100 13 A B FIG. S2: (A) The logarithm of the participation ratio P and of the norm of the shear pseudo-energy Esh is shown as function of the rank. The measured relative variation from the two curves is γ∗ (cid:39) 3.46 and we choose γ = 3.5 in Eq.S9. (B) Scatter plot of Γ versus the participation ratio P where Γ is averaged over the proteins with overlap larger than 0.45 and shown for (i) the modes (without considering the modes with maximal overlap), (ii) the modes with maximum overlap and (iii) the allosteric response. The different points correspond to different values of mode rank. This alternative visualization confirms the results presented in the main text. FIG. S3: (A) Face-centered cubic lattice with open boundaries. Harmonic springs of stiffness k are represented in clear blue and non-linear springs by red dashed lines. Black lines represent the links with stiffness k1 (cid:29) k, used to simulate the binding of a ligand. (B) A projection of the cubic lattice in the (x,z) plane. (C) A projection of the cubic lattice in the (x,y) together with the shear mechanism illustrated by the blue arrows . These examples are for a system with size N = 83. 020406080-1.2-1-0.8-0.6-0.4-0.2-6-5-4-3-2-1-0.8-0.6-0.4-6.5-6-5.5-5-4.5
1911.02526
1
1911
2019-11-01T16:18:23
Dynamic traversal of large gaps by insects and legged robots reveals a template
[ "physics.bio-ph", "eess.SY", "eess.SY", "q-bio.QM" ]
It is well known that animals can use neural and sensory feedback via vision, tactile sensing, and echolocation to negotiate obstacles. Similarly, most robots use deliberate or reactive planning to avoid obstacles, which relies on prior knowledge or high-fidelity sensing of the environment. However, during dynamic locomotion in complex, novel, 3-D terrains such as forest floor and building rubble, sensing and planning suffer bandwidth limitation and large noise and are sometimes even impossible. Here, we study rapid locomotion over a large gap, a simple, ubiquitous obstacle, to begin to discover general principles of dynamic traversal of large 3-D obstacles. We challenged the discoid cockroach and an open-loop six-legged robot to traverse a large gap of varying length. Both the animal and the robot could dynamically traverse a gap as large as 1 body length by bridging the gap with its head, but traversal probability decreased with gap length. Based on these observations, we developed a template that well captured body dynamics and quantitatively predicted traversal performance. Our template revealed that high approach speed, initial body pitch, and initial body pitch angular velocity facilitated dynamic traversal, and successfully predicted a new strategy of using body pitch control that increased the robot maximal traversal gap length by 50%. Our study established the first template of dynamic locomotion beyond planar surfaces and is an important step in expanding terradynamics into complex 3-D terrains.
physics.bio-ph
physics
Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu Dynamic traversal of large gaps by insects and legged robots reveals a template Sean W. Gart, Changxin Yan, Ratan Othayoth, Zhiyi Ren, and Chen Li Department of Mechanical Engineering, Johns Hopkins University 3400 N. Charles St, 126 Hackerman Hall, Baltimore, Maryland 21218-2683, USA E-mail: [email protected] Keywords: Locomotion, obstacle, cockroach, Blaberus discoidalis, bio-inspiration, terradynamics Abstract It is well known that animals can use neural and sensory feedback via vision, tactile sensing, and echolocation to negotiate obstacles. Similarly, most robots use deliberate or reactive planning to avoid obstacles, which relies on prior knowledge or high-fidelity sensing of the environment. However, during dynamic locomotion in complex, novel, 3-D terrains such as forest floor and building rubble, sensing and planning suffer bandwidth limitation and large noise and are sometimes even impossible. Here, we study rapid locomotion over a large gap, a simple, ubiquitous obstacle, to begin to discover general principles of dynamic traversal of large 3-D obstacles. We challenged the discoid cockroach and an open-loop six-legged robot to traverse a large gap of varying length. Both the animal and the robot could dynamically traverse a gap as large as 1 body length by bridging the gap with its head, but traversal probability decreased with gap length. Based on these observations, we developed a template that well captured body dynamics and quantitatively predicted traversal performance. Our template revealed that high approach speed, initial body pitch, and initial body pitch angular velocity facilitated dynamic traversal, and successfully predicted a new strategy of using body pitch control that increased the robot's maximal traversal gap length by 50%. Our study established the first template of dynamic locomotion beyond planar surfaces and is an important step in expanding terradynamics into complex 3-D terrains. 1 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu 1. Introduction Complex 3-D terrains such as leaf litter, fallen branches on a forest floor, and landslide debris (figure 1) can pose a major challenge for small animals and robots alike, because obstacles are often comparable to, or even larger than, the animal or robot itself [1] and can induce large perturbations [2 -- 8]. Despite this, small animals like insects, reptiles, and small birds agilely traverse complex 3-D terrains [2, 9] as diverse as inclined and vertical surfaces [10 -- 16], thin ledges [17], large gaps [4, 7, 18 -- 21] and bumps [3, 5, 22 -- 26], uneven surfaces [6, 27 -- 30], cluttered terrain [8], and even confined spaces [31], with locomotor performance far exceeding that of even the best robots today [32 -- 36]. To move through varying complex environments, animals not only use multi-modal sensory feedback and feedforward neural commands to control and adjust their body and limbs, but their body and limbs also often have well-tuned morphology to accommodate perturbations in the environment via mechanical feedback [2, 37]. During slow locomotion where sensing is sufficient, animals like stick insects [4, 38], cockroaches [3, 23, 39, 40], lizards [5, 24], and snakes [7, 19] tend to use deliberate, seemingly well-planned, and often precisely-controlled, body and limb movements to negotiate with and traverse complex 3-D terrains, presumably using antennae or vision in the process. For example, stick insects use their antennae to sense the terrain and use quasi-static, "follow-the-leader" stepping to slowly walk on branches [41] climb over steps [11, 42] and bridge over large gaps . Slowly running [39, 43] or walking [3, 23, 40, 44] cockroaches use their antennae to sense obstacles in front and alter their kinematics to either climb over steps [3, 22], tunnel under steps [23, 45], approach and climb up pillars [44, 46], or follow walls [39, 43, 47] depending on the location of the obstacle. Lizards frequently jump onto and over large bumps [48], and snakes either quasi-statically cantilever their body to reach across a smaller gap [19] or dynamically lunge to traverse a larger gap [19], presumably all using vision in the process. 2 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu Figure 1. Examples of gap-like obstacles in complex terrains: (a) Leaf and branch litter, (b) tree crevices, (c) cracked dry soil, and (d) building rubble. All images were available under Creative Commons CC0 at pixabay.com. By contrast, during rapid locomotion such as prey chasing and predator escape, particularly in terrains with frequent large disturbances [25], sensory noise [37] and sensory and neuromuscular delays [49] limit an animal's ability to plan its course of locomotion and control its body and limb movement during each step or gait cycle. To accommodate this, animals primarily move their body and limbs in a more feedforward manner [37] to simplify control, and use neural and sensory feedback to adjust body and limbs in response to large environmental perturbations. Because of this, how its body and limbs passively respond to the environment has a crucial impact on the dynamics and locomotor performance of a small animal moving dynamically through complex 3-D terrains [6, 8, 28, 50]. For example, cockroaches have streamlined body shapes that facilitate traversal of cluttered terrain , highly compressible yet robust exoskeletons that help them move through small crevices [31], sprawled leg posture that self-stabilizes lateral perturbations [51], viscoelastic leg muscles and tendons that dampen external perturbations [52, 53], and distributed leg spines that increase probability of firm foothold on low contact area surfaces [28]. 3 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu Traditionally, the field of mobile robotics has mainly dealt with locomotion in complex environments by solving the problem of obstacle avoidance [54 -- 56], which requires either prior knowledge or high-fidelity sensing to plan clear locomotion paths [56 -- 58]. More recently, thanks to cross-discipline collaborations between biologists, applied mathematicians, and engineers, the neuromechanical principles from biological studies have enabled many under-actuated bio-inspired robots [32, 35, 59 -- 61]. By combining high-level sensing and planning with mechanical feedback via mechanically tuned designs and control algorithms [62 -- 64], these robots have achieved unprecedented locomotion performance on simple ground [32, 59] and are beginning to traverse complex 3-D terrains [8, 31, 36, 61]. For both legged animals and robots, primary focus of mechanical feedback studies has been how leg morphology and mechanics interact with simple ground to assist locomotion [37, 63, 65]. In complex 3-D terrains where obstacles could be larger than the animal or robot, the body of a small legged animal or robot, which is more robust to collision [31, 50, 66] may also physically contact and interact with the terrain to help locomotion. However, much less attention has been paid to the role of body-terrain interaction in legged locomotion until very recently [31, 50]. Here, we take the next step in understanding how body-terrain interaction affects dynamic traversal of large 3-D obstacles. We chose to focus on two simple large obstacles: (1) a gap obstacle as large as the animal / robot's body length, reported in this paper; and (2) a bump obstacle as high as four times the animal / robot's hip height, reported in a companion paper [67]. Such simple, well-defined, parameterizable laboratory models of natural terrains are useful towards understanding complex interactions between moving animals and their natural environment [3, 8, 10, 14, 68]. We chose to study the discoid cockroach, which lives on the floor of tropical rainforests and naturally traverses a wide variety of 3-D obstacles such as dense vegetation and litter [69], because mechanical feedback plays a major role in its locomotion at high speeds [6, 28, 37]. Better ability to traverse large gap and bump obstacles is also important for small legged robots in complex 3-D terrains like building rubble and landslides (figure 1(c)) during applications like search and rescue [70]. We chose to study a cockroach-inspired RHex-like robot [32] because it not 4 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu only has similar body plan and running dynamics [71] as the animal and allows a direct comparison with animal observations, but also provides a physical model to precisely control and systematically vary locomotor parameters [70, 72]. In this paper, we focus on understanding how legged animal and robot's body dynamics and body- terrain interaction affects dynamic traversal of a large gap obstacle. We challenged the rapidly running animal and the open-loop robot with a large gap of variable size, and tested how the ability to traverse depended on gap length, running speeds towards the gap, and body orientation. Comparison of the animal with the open-loop robot provided insights into the role of active sensory feedback. Inspired by the similarities of animal and robot observations, we developed a reduced-order dynamic model, or template [10, 37, 73], to capture low-order dynamic traversal of a large gap obstacle. Finally, we added an active tail [21, 74] to the robot and tested a control strategy revealed by our template to enhance large gap traversal. In our companion paper [67], we report our experiments and modeling of dynamic traversal of a large bump obstacle, and discuss common and potentially general, principles and distinct differences between dynamic traversal of large gap and bump obstacles. 2. Methods 2.1. Animals For animal experiments, we used seven male Blaberus discoidalis cockroaches (Pinellas County Reptiles, St Petersburg, FL, USA), as females were often gravid and under different load bearing conditions. Prior to experiments we kept the animals in individual plastic containers at room temperature (22 °C) on a 12h:12h light:dark cycle and added food (fish and rabbit pellets) ad libitum. Animals weighed 2.7 ± 0.2 g and measured 4.6 ± 0.2 cm in length, 2.3 ± 0.2 cm in width, and 0.7 ± 0.1 cm in thickness. 2.2. Legged robot 5 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu For robot experiments, we constructed a cockroach-inspired, six-legged robot by modifying the RHex-class design37 (figure 2(b)). The slightly-flexible robot chassis was cut from a 3.1 mm thick acrylic sheet using a VLS 6.60 laser cutter (Universal Laser Systems Inc., Scottsdale, AZ). We attached each motor (460 RPM micro-gear DC motor, 50:1 gear ratio, ServoCity, Winfield, KS) that drove the legs to 2 cm × 2 cm pieces of 0.15 cm thick polystyrene that were subsequently attached to the chassis. To increase the maximal leg frequency given the motor used to drive the legs, we chose s-shape legs. The s-shape legs were 3D printed with PLA plastic (Ultimaker 2 Extended +, Ultimaker North America, Cambridge, MA). We wrapped each of the legs with friction tape (Duck Brand, Avon, OH) to increase ground traction. To generate stable spring-mass-like [75] running, we tuned the stiffness of the chassis and legs and leg friction. To approximate the anterior shape of the cockroach body for direct comparison of terrain interaction between the robot and the animal, we thermo-formed a bottom half of a quarter ellipsoid polystyrene shell, and attached it to the front of the robot. The robot measured 25 cm long, 16 cm wide, 8 cm tall, and weighed 194 g. We calculated robot moment of inertia by approximating the body as a simple rigid rod measuring 25 cm long and weighing 194 g, which rotates about a fixed end (I = 1/3mb2 = 0.004 kg m2, where m is body mass and b is body length). To test the effect of passive mechanics, we did not implement any sensors and drove the robot with an open-loop leg control. We varied the robot's running speed from 50 cm s -- 1 to 200 cm s -- 1 by changing voltage supplied to the DC motors to adjust leg frequency. At the maximal voltage of 25 V, the robot ran stably at 8.0 ± 0.5 body lengths s -- 1 (200 ± 12 cm s -- 1). 2.3. Gap obstacle track We constructed a 90 cm long, 30 cm wide track with a gap that spanned the entire width of the track (figure 2(c)) using t-slotted extruded aluminum and acrylic (McMaster-Carr, Elmhurst, IL, USA). For both animal and robot experiments, we varied gap length L from 0.2 body length to 1 body length (b ≈ 5 cm for the animal, b = 25 cm for the robot) by sliding the far side of the track. To measure the maximal traversable gap length, we made an infinite length gap (or cliff) by removing the far side of the gap. We 6 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu covered the entire test surface with white paper cardstock (Pacon 4-ply railroad poster board, Appleton, WI, USA) for animal experiments. For robot experiments, we covered the near side of the test surface with 50 grit sandpaper to increase leg traction and prevent slipping when the legs started moving. To prevent scratching of the robot shell during impact, we covered the opposite side of the gap with polystyrene. 2.4. Experiment protocol We filmed the animal and robot running over the gap at 500 frames s−1 using three synchronized high-speed cameras, two from the dorsal view and one from the side view (figure 2(c)), with a shutter time of 500 µs. Dorsal cameras were placed directly above the near edge of the gap. A small lens aperture was used to maximize the focal depth of field. We illuminated and heated the test area to 35 °C with 500 W work lamps (Coleman Cable, Waukegan, IL, USA), three from the dorsal side, and two laterally. To track the animal and robot, we attached a BEEtag [76], printed on standard office paper and attached to cardstock using double-sided tape, to the dorsal side of the body. 7 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu Figure 2. Experimental setup and definition of geometric and kinematic parameters and variables. (a) Discoid cockroach, body length b = 4.6 cm ± 0.2 cm, hip height zhip = 0.5 cm. (b) Legged robot, body length b = 25 cm, hip height zhip = 2.5 cm. BEEtags [76] were used to measure 3-D position (x, y, z), orientation (yaw α, pitch β, and roll γ), and pitch angular velocity ω of the body. (c) Schematic of the gap obstacle. Gap length L was varied from 0.2 to 1 body length for both animal and robot experiments. Gap depth was fixed (D = 0.7 body length for animal; D = 0.3 body length for robot). Two dorsal and one lateral high speed cameras were used to record locomotion. 2.4.1. Animal experiments We filmed animal experiments using Photron Mini UX100 cameras (Photron USA, San Diego, CA, USA) with a resolution of 1280 × 1024 pixels. For animal tracking, we attached a 1.6 cm × 1.6 cm BEEtag [76] to the dorsal surface of the wings directly above the body center of mass (CoM) [77] (figure 2(a)). We attached the tag to the animal using a combination of hot glue, super glue, and baking powder (as an accelerant). The animals were allowed to rest for at least 1 hour after tags were attached. 8 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu We placed cockroaches on the track one at a time for testing. To elicit a rapid escape response, we prodded the posterior and abdomen of the animal with a rod wrapped in paper tape. The animal ran towards the obstacle between two walls that funneled it towards the middle of the track. To encourage the animal to seek shelter [23, 45, 78], we placed a shaded overhang after the obstacle. The animals were allowed to rest for 1 to 2 minutes before each trial. 2.4.2. Robot experiments We filmed all robot experiments using Fastec IL5 cameras (Fastec Imaging, San Diego, CA, USA) with a resolution of 1920 × 1080 pixels. For robot tracking with no tail, we attached a 5 cm × 5 cm tag above the robot body CoM using 0.3 cm thick polystyrene. The polystyrene was robust in cases where the robot flipped over. For experiments with a tail, we placed an additional 5 cm × 5 cm tag on the front shell of the robot, and the body CoM tag was moved to the side of the tail servo motor so that it was not obstructed by the tail when actuated. Three 3.8 cm × 3.8 cm tags were added to the dorsal side, ventral side, and tip of the tail to track tail motion. 2.4.3. Tailed robot experiment To study how initial body pitch and initial body pitch angular velocity affected gap traversal, we added an active tail to the posterior end of the robot. We fastened the base of the 12 cm long active tail (3- D printed from PLA plastic) to a high torque servo motor (Futaba BLS274SV, Futaba, Champaign, IL) and attached a 33 g mass at its distal end. The active tail rotated within the sagittal plane at a maximal angular speed of 315 ± 115 ° s−1 . We tuned the base position (6 cm from the posterior end) and actuation timing (140 ± 14 ms prior to reaching the obstacle) of the active tail so that the robot's body pitch increased from −4° ± 8° prior to actuation to 8° ± 6° after actuation. The servo motor actuation was controlled by an Arduino Uno micro-controller and a motor controller (Qunqi L298 Dual H-bridge motor driver module). For this experiment, the robot legs were still under open-loop control. Adding the tail increased the robot's 9 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu total body inertia along the pitch axis about its posterior end from 0.004 kg m2 to 0.007 kg m2 and did not change the robot's maximal running speed (P = 0.4, Student's t-test). To test the effect of body pitching, we ran the tailed robot perpendicular to the gap at a constant speed of 190 ± 16 cm s -- 1 and varied whether the tail was activated or not for each gap length. We empirically timed tail activation so that the robot started to pitch up as the head reached the near edge of the gap. For finite length gap experiments, we increased gap length until the robot failed to traverse in all trials and decreased it until the robot successfully traversed in all trials. We collected 10 trials for each gap length tested. This resulted in a total of 40 trials for the experiment without tail actuation for a gap of 0.6, 0.8, 1.0, and infinite body length, and a total of 50 trials for the experiment with tail actuation for a gap of 0.8, 1.0, 1.2 1.4, and infinite body length (N = 1, n = 90). 2.4.4. Kinematic Tracking To calibrate the cameras, we made a calibration grid with Lego bricks (The Lego Group, Bilund, Denmark) and placed it in view of all cameras prior to each experiment session. We obtained intrinsic (focal length and lens distortion) and extrinsic (relative positions and rotations) camera parameters using direct linear transformation (DLT) 3-D reconstruction [79]. After experiments, we imported image sequences into a custom MATLAB script that tracked the tags in each camera view using the BEEtag code [76]. With an additional custom DLT script [79], we obtained the 3-D position (x, y, z) and orientation (Euler angles yaw α, pitch β, and roll γ) of the tags (figures 2(a, b)). To verify BEEtag accuracy, we 3-D printed a calibration object with 9 BEEtags equally spaced 7 cm apart in a grid in the horizontal plane, but orientated at pitch and yaw angles from 0° to 60° in an increment of 30°. We then calculated error by measuring the 3-D position and orientation all the tags and comparing them with the designed values. We found that BEEtags accurately measured body orientation angles (s.d. of error = 1.1°) and fore-aft position and body height (s.d. of error = 1.2 mm). We measured the position and orientation of the CoM relative to the tag and inferred the CoM for each animal. To examine how the animal and robot's head interacted with the obstacle, we calculated head position from the tag position, assuming that the body and head together acted as a rigid 10 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu body. This was a reasonable approximation: our manual tracking verified that the head moved little (< 0.13 cm) relative to the tag throughout each running cycle, and only moved slightly longer (0.27 cm) when the animal flexed its neck joint and abdomen while gripping and climbing onto the far side of the gap. For the robot, we generated a 3-D point cloud of the shell using a CAD model (Solidworks, Solidworks corporation, Waltham, MA, USA) to obtain head position (anterior end of the shell) relative to tag position. 2.5. Data analysis For all animal and robot experiments, we measured 3-D kinematics including speed v, body pitch β, body pitch angular velocity ω, body yaw α, and body roll γ as a function of forward displacement x and traversal probability. We categorized a trial as successful traversal if the animal and robot's CoM reached the far side of the gap, and as failure to traverse if any part of the body touched the bottom of the gap. For the infinite length gap experiment, we also measured the distance that the animal and robot's head reached before it fell below the sides of the gap (maximal traversable gap length, d). For all experiments, we observed only a small difference between the velocity heading and body yaw immediately prior to gap encounter (7° ± 8° for the animal, 2° ± 8° for the robot). Therefore, we assumed that velocity heading always equaled body yaw. We defined angle of incidence θ0 as the angle between velocity heading and the forward x direction at the time of obstacle encounter, which equaled body yaw 0 at the time of obstacle encounter (see figure 2(a, b)). We calculated angle of incidence 0, initial body pitch β0, initial body pitch angular velocity ω0, initial body roll γ0, and approach speed v0=vcosθ0 (speed perpendicular to the gap) immediately before the animal and robot encountered the near edge of the gap obstacle. Our use of approach speed accounted for any motion not perpendicular to the gap. All metrics were averaged over 4 frames to reduce instantaneous measurement error, except for initial body pitch angular velocity averaged from when the head reached until the body CoM passed the near edge of the gap to account for the large noise. We reported both body yaw  and angle of incidence 0 in absolute values due to lateral symmetry. 11 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu To determine whether the animal and robot approached and traversed the gap obstacle at speeds comparable to that during walking or running, we used the Froude number [80] Fr = v0(gzhip)−1/2, where g is acceleration due to gravity, v0 is approach speed, and zhip is hip height (0.5 cm for the animal, 2.5 cm for the robot). As a common form of normalized speed of terrestrial locomotion, Froude number is a good predictor of the speed at which legged animals transition from a walking to a running gait [80]. For the discoid cockroach, this transition occurs at normalized speeds of Fr = 1.5 to 1.7 [6, 71]. In the remainder the paper, we refer the locomotion speed perpendicular to the gap as approach speed. 2.6. Statistics For animal gap experiments, seven animals ran 10 trials each over six different gap lengths of 0.2, 0.4, 0.6, 0.8, 1, and infinite body length, resulting in a total of 420 accepted trials (N = 7, n = 420). We rejected a trial whenever the animal collided with the wall, turned back, or stopped moving forward before encountering the gap, or if it moved out of the camera view during a traversal attempt. We varied gap length but allowed the animal to run with its own chosen speed and velocity heading during each trial. For the robot gap experiment, we systematically varied gap length from 0.2 to 1 body length in increments of 0.2 body length and tested five different running speeds of 78 ± 15 cm s−1, 120 ± 17 cm s−1, 170 ± 18 cm s−1, 190 ± 10 cm s−1, and 200 ± 12 cm s− 1. The robot initial velocity heading was always perpendicular to the gap. We collected 5 trials for each combination of gap length and speed, resulting in a total of 110 trials (n = 1, N = 110). For the infinite length gap experiment, we tested five different approach speeds of 56 ± 13 cm s−1, 78 ± 3.0 cm s−1, 130 ± 11 cm s−1, 180 ± 7.6 cm s−1, and 200 ± 6.2 cm s−1 and collected 5 trials at each speed for a total of 25 trials (n = 1, N = 25). With no tail, the robot had a small, nearly constant initial body pitch of 1° ± 3° (mean ± s.d.) immediately before reaching the near edge of the gap. To compare metrics (approach speed, traversal probability, etc.) across a gap of different lengths, we first calculated the mean for each individual for each gap, and then averaged over individual means to 12 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu obtain the cross-individual average for that gap. Because we used only one robot, we averaged all trials for each gap to obtain the mean. To test which metrics (approach speed, initial body pitch, etc.) affected traversal success or failure, we used multiple logistic regression. To test all other metrics reported, we used repeated-measures ANOVA for animal experiments and ANOVA for robot experiments. Our multiple logistic regression and repeated-measures ANOVA accounted for individual variance by using the individual as one of the factors. We used Tukey's honestly significant difference (HSD) method for post- hoc analysis where needed. All data are reported as means ± 1 standard deviation (s.d.) unless otherwise specified. 3. Results and discussion 3.1. Dynamic gap traversal performance In our experiments, both the discoid cockroach and the robot dynamically ran at high speeds to traverse the gap obstacle. Animal running speeds were 61 ± 16 cm s−1 (12 ± 4 body length s−1), up to 9 times that in previous experiments of slow, quasi-static gap traversal [3, 16, 23]. The majority (83%) of animal trials had a Froude number above that for walking-to-running transition (Fr = 1.5) [6, 71]. Robot running speeds were 146 ± 48 m/s (6 ± 2 body length s−1). During such high speed locomotion, both the animal and the robot were able to dynamically traverse a large gap of up to 1 body length. Probability of dynamic traversal decreased with gap length (figure 5(a, b), solid curve; animal: P < 0.0001, multiple logistic regression; robot: P < 0.0001, multiple logistic regression). When the gap is small enough (up to 0.4 body length), both the animal and the robot almost always traversed (93% for the animal and 100% for the robot, respectively). By contrast, for the largest gap of 1 body length tested, traversal was unlikely for both the animal and the robot. Only two out of the seven cockroaches were able to traverse it, while the rest always failed to traverse; the robot was 13 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu never able to traverse it at its highest running speed (8.0 ± 0.5 body lengths s−1). We verified that no animals could traverse a gap larger than 1 body length. 3.2. Gap bridging by head well predicts dynamic gap traversal Careful observation of how the animal and robot's body interacted with the gap revealed that dynamic traversal of a gap obstacle was well predicted by whether the head was able to bridge across the gap. As the animal and robot encountered a large gap, each pair of legs sequentially lost contact with the near side, and the body began falling, with the head pitching downward (figure 3(a, b, e, f), frames 1, 2, and 3; supplementary video 1, 2). Gap bridging by the head occurred when the animal or robot's head reached the far edge of the gap before falling below it (figure 3(a, e); figure 3(c, d, g, h), red solid curve). When gap bridging by the head did not occur, the animal and robot always fell into the gap and failed to traverse (figure 3(b, f); figure 3(c, d, g, h), blue dashed curves). We found that when gap bridging by the head occurred, traversal was successful 97% of the time for the animal and 81% of the time for the robot (the exceptional cases were due to failure to grip, discussed in Section 3.6). When the head did not bridge the gap, both the animal and the robot failed to traverse 100% of the time. For a small enough gap (below 0.6 body length), the animal and robot always traversed because the head could not fall below the surface over such a small forward displacement (figure S4, S5), and gap bridging was guaranteed. 14 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu Figure 4. Dynamic locomotion of the discoid cockroach and the robot over a large gap obstacle. Representative trials of (a, e) successful traversal and (b, f) failure to bridge the gap. (c, g) Head height as a function of forward position of the head.* (d, h) Body pitch as a function of forward position of the head. In (c, d, g, h), solid red and blue dashed curves and shaded areas represent means ± 1 s.d. for the cases of success and failure. Data are shown for the 0.8 body length (4 cm) gap as an example; data for other gap lengths have similar trends (figures S4, S5, S6, S7). For simplicity, only movement perpendicular to the gap (within the x-z plane) is shown. * We noted that when the animal failed to traverse, it often rebounded backwards after impacting the far side of the gap, but we could not track this motion due to the body and gap ledge obscuring the tag. Therefore, we truncated the data at the far edge of the gap. 15 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu The animal's gap bridging using the head was likely a passive process, similar to our robot under open-loop control. Although cockroaches actively sweep their antenna to sense the physical environment during slow exploration [40], we observed that the animal's antenna were held straight forward and slightly upward [40] and rarely came in contact with the ground. In addition, at the high speeds in our experiments, the animal had a very short time to respond if it did detect the gap (given a sensory delay of 6-40 ms [45] and a neuromuscular delay of 47 ms [6]), comparable to the time (60 ± 10 ms) for the animal's head to fall and bridge across the largest gap. Further, we observed that the animal rarely changed speed (by 3 ± 4 cm s−1), body pitch (by 2° ± 3°), or heading (by 6° ± 6°) before reaching the gap. Together, these observations indicated that the animal was likely unable to respond in time. By contrast, later during dynamic gap traversal, the animal likely used sensory feedback to initiate and control active body and leg adjustments (see section 3.6). 3.3. Template for dynamic gap traversal The striking similarities in traversal performance of the animal and the robot and the signature observation of gap bridging by the head for successful traversal suggested a template for dynamic gap traversal. A template is a useful modeling concept that allows general, fundamental understanding of high- dimensional, nonlinear, multi-body dynamic locomotion phenomena by reducing the problem to as few degrees of freedom as possible [37, 75]. A few templates have already captured fundamental dynamics and revealed control strategies for common forms of terrestrial locomotion on 2-D surfaces such as dynamic walking [81, 82] and running on level ground [62, 63, 83] and dynamic climbing on vertical walls [10, 15]. Inspired by these successes, we take the next step in creating a template for dynamic traversal of a large gap as a representative of complex 3-D terrains. We approximated the animal and robot body as a rigid ellipsoid traveling forward at a constant approach speed, and calculated its dynamics during passive falling under gravity as it encountered a large gap (figure 4(a)). For simplicity, we assumed that the rigid body only rotated about the body pitch axis in the sagittal plane and had no body yaw or roll rotations. In addition, we assumed a fixed axis of rotation at 16 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu the posterior end of the body at a height equaling the hip height of the animal (0.5 cm) or robot (2.5 cm). Finally, to model the body gradually losing support as legs gradually lost contact when the animal or robot ran past the near edge, we assumed that the force due to gravity increased proportionally to the length of the body beyond the near edge. Using the Lagrangian method, we obtained the equation of motion during the pre-free falling phase (before the posterior passed the near edge) as: d dt (∂L ∂ω )  ∂L ∂β =0 (1) with the Lagrangian L= 1 2 Iω2  mgb 2 sin β (2) where β is body pitch, I = 1/3 mb2 is the moment of inertia for a body pitching about the hinge on its end, b is the length of the body (5 cm for the animal, 25 cm for the robot), m is body mass (2.5 g for the animal, 194 g for the robot), ω is the pitch angular velocity, and g = 9.81 m s−2 is the gravitational acceleration. From the equation of motion, we obtained angular acceleration about the body pitch axis: ω (t)= ξmgb 2I cos β(t) (3) where ξ= (vt b ) cos β(t) cos θ0 is a linear force reduction factor that characterizes the loss of body support due to the legs losing surface contact, v is the speed along the long axis of the body, and θ0 is the angle of incidence that measures the deviation of body yaw and velocity heading from the forward direction (direction perpendicular to the gap). Once the posterior end of the body reached the near edge of the gap, the body started to fall like a projectile with zero angular acceleration and ω(t) = ωedge, where ωedge is the angular velocity of the body when the posterior reached the near gap edge. The center of mass height during the free falling phase is: 17 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu zCoM(t) = 1 2 gt2 , (6) ω (t) = 0 (7) where t is the time after the posterior end passed the near edge of the gap. Note that we assumed that velocity heading and body yaw were always aligned based on experimental observations (see Section 2.5). Finally, we used the Euler method to integrate forward in time to obtain body pitch angular velocity ω, body pitch β, the center of mass position xCoM along the forward direction, and the vertical center of mass position zCoM as a function time t: ω(t+Δt) = ω(t) + ω (t)Δt (4) β(t+Δt) =β(t)+ β(t)Δt (5) xCoM(𝑡+Δt) =xCoM(𝑡)+𝑥 CoM(𝑡)Δt zCoM(𝑡+Δt) = zCoM(𝑡)+ 𝑧CoM(𝑡)Δt (6) (7) where time step Δt = 0.001 seconds. Assuming that gap bridging by the head results in traversal, the template allowed us to predict the maximal traversable gap length that the body could traverse for any given initial body pitch β0, initial body pitch angular velocity ω0, and approach speed v0 (supplementary video 4). We first numerically calculated time that the body had to attempt to bridge the gap, tbridge, i.e., the time for the rigid body's head (free anterior end) to fall to surface level (zero height). The forward distance over which the head travelled, d = v0tfall, was then the maximal traversable gap length (figure 4(b)). 18 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu Figure 4. Template for dynamic gap traversal. (a) An animal or robot's body is modeled as a rigid body traveling at a constant speed v along its body long axis, with a body pitch angle β and body pitch angular velocity . As it encounters a gap at an angle of incidence , the body rotates downward in the sagittal plane about a hinge at its posterior end with a height equaling the animal or robot's hip height zhipThe body yaw and velocity heading are assumed to be the same and at an angle  relative to the direction perpendicular to the gap. (b) Side view of model-predicted representative head trajectories and maximal traversable gap length, d, given the initial conditions ω0 = 0° s−1, β0 = 20°, and a higher speed v0 = 65 cm s−1 (red solid curve), bridging the gap, and a lower speed v0 = 30 cm s−1 (blue dashed curve), falling into the gap. To validate the template, we ran the animal and the robot into an infinite gap (a cliff) to measure maximal traversable gap length and compared it with model predictions. Without any model fitting parameters, the template well predicted maximal traversable gap length d for both the animal and the robot (figure 5(a, b)). We noted that the animal and the robot ran at a broad range of approach speed (animal: 37 cm s−1 ≤ v0 ≤ 100 cm s−1; robot: 39 cm s−1 ≤ v0 ≤ 210 cm s−1), initial body pitch (animal: 3° ≤ β0 ≤ 24°; robot: 19 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu −5° ≤ β0 ≤ 9°), and initial body pitch angular velocity (animal: −401° s−1 ≤ β0 ≤ 34° s−1; robot: −143° s−1 ≤ β0 ≤ 103° s−1) in the experiment. The template slightly under-predicted maximal traversable gap length for the robot. One reason was due to its bouncier running gait, which resulted in a significant initial upward speed (20 ± 20 % of its forward speed, vs. the animal's 5 ± 4 %), not accounted for by the horizontally approaching template. Additionally, because the robot's head was in front of its legs by a large distance (0.3 body length), it did not start to lose ground reaction force and fall into the gap until it had already entered the gap substantially. By contrast, in the model, the body started to lose ground reaction force and fall immediately upon reaching the gap. Next, we tested the predictive power of the template for predicting traversal probability measured in finite length gap experiments. Using the measured approach speed v0 = vcosθ0, initial body pitch β0, and initial body pitch angular velocity ω0 from each experimental trial as initial conditions, we calculated the expected maximal traversable gap length from the model. If it exceeded the gap length being tested, the model-predicted successful traversal (figure 4(b), red solid curve); otherwise, the model-predicted failure (figure 4(b), blue dashed curve). Without any model fitting parameters, the template well predicted the observed monotonic decrease of traversal probability with gap length for both the animal and the robot (figure 5(c, d), dashed curves). For small gap length, quantitative agreement between animal data and model predictions was excellent. The over-prediction of traversal probability for the two largest gaps was because the model did not account for the animal or robot failing to grip the far side of the gap after the head bridged (see section 3.6). The template under-predicted robot traversal probability for the 0.4 and 0.6 body length gaps due to its bouncier gait and its leg substantially behind the head discussed above. 20 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu Figure 5. The template well predicted dynamic gap traversal. (a, b) Comparison of maximal traversable gap length measured from the infinite length gap experiment with that predicted from the model. The black diagonal line indicates perfect match between data and model. (c, d), Comparison of experimental (filled circles, solid curve) and model-predicted (open circles, dashed curve) traversal probability. Error bars represent 95% confidence interval. 3.4. Principles of dynamic gap traversal from template Our experimentally validated template allowed us to gain insights into general principles of dynamic gap traversal. Using the template, we numerically calculated how maximal traversable gap length depended on approach speed v0, initial body pitch , and initial body pitch angular velocity  over a broad range of parameter space. We discovered that maximal traversable gap length increased with all these 21 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu locomotor parameters (figure 6). This is intuitive because faster running allows an animal or robot to travel forward by a larger distance before its head falls below the surface level, and higher initial body pitch and higher initial pitch angular velocity gave the animal or robot longer time to travel forward before the head fell below surface level. These general principles from the template gave us two predictions: First, for a given gap length, the animal or robot running at higher approach speed, higher initial body pitch, and/or higher initial body pitch angular velocity should be more likely to successfully traverse. Second, as gap length increased, the approach speed, initial body pitch, and initial body pitch angular velocity required to traverse should also increase (supplementary video 4). Indeed, in our finite length gap experiments, we observed that for all but the smallest gap lengths tested, both the animal and robot ran at higher average approach speeds (measured in Froude number, Fr) and had a higher initial body pitch angular velocity when they traversed the gap compared to when they failed (figure 6, circles; figure A2; animal: P < 0.05, multiple logistic regression; robot: P < 0.05, multiple logistic regression). However, for both the animal and the robot, initial body pitch did not differ between successful and failed trials for all the gap lengths tested (P > 0.05, multiple logistic regression). This was likely a result of the small range and large variation of initial body pitch attainable by the animal (12° ± 4°) and the robot (1° ± 3°) when they ran at the speeds in our experiments. In the infinite length gap experiments, both the animal and robot reached a maximal traversable gap length longer than its body length (figure 5(a, b); animal: 1.25 body length; robot: 1.6 body length). However, traversal of these gap lengths are unlikely. The animal would collide with the far side of the gap at high speeds, making gripping difficult due to the slow reaction time. Additionally, the animal and the robot often could not grip before falling into the gap. Therefore, the model likely predicted a larger maximal traversable gap length than is physically possible due to the grip failure (see section 3.6). 22 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu Figure 6. The template quantitatively predicted maximal traversable gap length for both the animal and the robot. Maximal traversable gap length as a function of approach speed (shown in Froude number Fr = v0(gzhip)1/2) and initial body pitch, and initial body pitch angular velocity for the animal (a) and robot (b). The surface colors show the model predictions. To create the contour slices, we used the initial body pitch angular velocity (horizontal plane) and the initial body pitch (vertical plane) averaged for all trials for each individual animal and for all robot trials. Circles show the mean ± s.d. for successful (red) and failed (blue) traversal of the 1 body length gap. For the tailed robotfilled circles show the mean ± s.d. without tail activation and open circles show the mean ± s.d. with tail activation (see section 3.5). Bracket and asterisk indicate a statistically significant difference. Given its simplicity and predictive power of dynamic gap traversal over a broad range of locomotor and terrain parameter space, our model provided the first template (figure 7) for dynamic locomotion in 23 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu complex 3-D terrains. In addition, because the model did not include any information of legs, we expect that the template can be applied to other types of locomotors, such as wheeled and tracked vehicles [84, 85] and limbless animals [86] and robots [87 -- 89] during high-speed, dynamic gap traversal. Figure 7. Summary of animal, robot, and template body dynamics and body-terrain interaction. Red and blue tubes and curves represent means ± 1 s.d. of vertical and lateral position of the head as the body moved forward towards the gap. The trajectories of successful traversal and failure are slightly offset in the y- direction for clarity. For simplicity, only movement perpendicular to the gap (within the x-z plane) is shown. 3.5. Body pitch control to enhance dynamic gap traversal The principles from our template provided useful control strategies for robots to traverse large gap obstacles. For example, for a given robot already reaching its maximal speed with a head-on approach (zero angle of incidence ), higher initial body pitch and higher initial body pitch angular velocity would allow it to traverse a larger gap. To demonstrate this, we tested the robot with a bio-inspired active tail [21, 33] that enabled adjustment of body pitching (figure 8(a)). As the robot approached the gap, the tail was rotated backwards and suddenly stopped, and its angular momentum was quickly transferred to the body, causing the robot to pitch upward (supplementary video 3). 24 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu As predicted by our template, we found that the active tail significantly increased the robot's ability to traverse a large gap by increasing 0 and the robot's pitch moment of inertia (figure 9). At an approach speed of 190 cm s−1 ± 20 cm s−1, the robot's initial body pitch angular velocity increased from -- 11 ± 74 ° s−1 without tail actuation to 57 ± 57 ° s−1 with tail actuation (figure 6(d), circles; P = 0.011, ANOVA). Although tail actuation did not significantly increase initial body pitch (P = 0.8, Student's t-test), it did increase the maximal body pitch as the robot ran over the near edge of the gap, from 5° ± 6° without tail actuation to 8° ± 6° with tail actuation (P = 0.045, ANOVA). Together, these changes in body pitching not only increased the robot's traversal probability for all but the smallest gap tested, but also allowed it to traverse a gap as large as 1.2 body length, a 50% increase (figure 9(b), open circles). Although we only demonstrated using an active tail to increase initial body pitch and initial body pitch angular velocity to aid large gap obstacle traversal, body pitch can continue to be controlled during the entire falling phase [21, 74]. Further, other body pitch control methods may also be employed, such as pushing more forcefully with fore and mid legs [34, 36] and hyperextending a flexible body [90]. Future studies should test the feasibility and performance of these control strategies and add sensory feedback [2] to further improve robot performance traversing large gap obstacles. Figure 8. Tailed robot experiments to demonstrate the novel control strategy for dynamic gap traversal. (a) The robot with an active tail. (b) Gap traversal probability for the tailed robot with (open circles) and without (filled circles) tail actuation while running at the same speed of 190 ± 20 cm s−1. 25 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu 3.6. Body flexibility and leg gripping ability enhance large gap traversal Our template of dynamic large gap traversal well captured body dynamics until the head bridged across the gap using kinetic energy. However, for traversal, the animal or robot must also be able to continue to move forward until its entire body made it across the gap. This is especially challenging over a large gap, because when the head bridged, the majority of the body or even the entire body is still above the gap without ground reaction forces (figure 3(a, b, c, d), frame 2). Therefore, generating sufficient upward and forward forces against the far edge of the gap is essential for traversal. Careful examination revealed that successful traversal of a large gap obstacle after gap bridging by the head often required additional leg gripping and pulling (66% and 100% of the time for 0.8 and 1 body length gaps, respectively), presumably using sensory feedback (figure 9(a)). The animal's ability to flex its body and use the various structures on its articulated legs and feet [91 -- 93] to grip when they touched the far side surface played an important role in this process. We observed that, after the animal's head bridged across a large gap, the body often flexed substantially, while its middle and fore legs pulled on the far edge and its hind legs pushed against the vertical surface (figure 9(a), frames 3, 4, 5). Body flexion not only allowed fore legs to better reach forward and downward to pull but also allowed hind legs to reach downward and forward to push [90]. When the animal was unable to do these sufficiently in time, its body pitched backwards, resulting in falling back into the gap even when the head bridged across (figure 9(b), frames 3, 4, and 5). By contrast, the robot the robot's body and legs are relatively rigid and lacked gripping mechanisms to pull itself onto the far side of gap (figure 9(c)). Even when the robot succeeded in traversing a large gap, it did not grip, but simply continued to use the same open-loop gait. As a result, even when the head bridged across the gap, the robot was more likely (19% probability) to fall backwards and fall back into the gap than the animal (3% probability) (P < 0.0001, repeated-measures ANOVA). In addition, the probability of grip failure for both the animal and robot increased for the largest gap lengths (animal: P < 0.0001, repeated- measures ANOVA; robot: P < 0.0001, ANOVA). 26 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu Based on these observations, we posit that a two-link body modeling the flexible body with three spring legs modeling fore, mid, and hind feet is a likely candidate for an anchor-level model [37] to better capture the dynamics of the final phase of dynamic traversal of large gaps and to further explore the role of active body pitch control [21] (figure 9(d)). Future experiments should better understand the biological principles of such active body and leg use and validate the anchor model, and use them to improve the ability of robots to dynamically traverse large gaps. Figure 9. Traversal often required gripping and body flexion. (a) Successful leg gripping aided by body flexion by the animal when dynamically traversing a large gap. (b) Grip failure of the animal. (c) Grip failure of the robot. (d) A proposed two-link anchor model for dynamic gap traversal with active body flexion and leg gripping. 27 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu 4. Conclusions We comparatively studied rapid-running discoid cockroaches and a cockroach-inspired robot as a physical model to discover the performance limits and general principles of dynamic traversal of a large gap obstacle. We discovered that similar to bridging small gaps between footholds on low contact area surfaces [28] and uneven terrain [6], both the animal and the robot can use translational and rotational kinetic energy to dynamically traverse a gap obstacle as large as its body length. This is rarely possible during quasi-static gap traversal [4, 38, 94] and more similar to snakes using kinetic energy to lunge across large gaps [19]. Traversal was less likely as the gap became larger and was well predicted by whether the head bridged across the gap. Inspired by the similarity in animal and robot observations, we created a template that well described body dynamics during passive falling over the gap and quantitatively predicted traversal performance. Our template revealed that high approach speed, high initial high body pitch, and high initial body pitch angular velocity all facilitated dynamic traversal, by allowing the head to travel further to bridge a gap larger than would be possible during quasi-static gap traversal [4, 38, 94]. Despite their similarities, the animal is still far better than the robot at dynamically traversing large gap obstacles, thanks to its stronger ability to grip, push, and pull itself onto the far side of the gap. Our study is a major step in expanding the emerging field of terradynamics of biological and robotic locomotion [8]. Our template is the first to quantitatively predict dynamic locomotion beyond planar surfaces [10, 37, 43] and expanded the usefulness of templates [37] into complex 3-D terrains. Future studies to systematically vary locomotor and terrain parameters [8, 72, 95, 96] and create new templates for other types of terrains will advance our understanding of how animals move in nature and improve robotic locomotion in complex natural and artificial environments. An immediate next step is to discover general principles for dynamic traversal of another simple yet general obstacle, a large bump, which we explore in our companion study [67]. 28 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu Acknowledgements We thank Yuanfeng Han, Tom Libby, Simon Sponberg, Bob Full, and two anonymous reviewers for helpful suggestions; Nastasia Winney and Rafael de la Tijera Obert for help with preliminary experiments; Yuanfeng Han for help with experimental setup; and Nicole Mazouch and Nate Hunt for providing animals during preliminary experiments. This work is funded by a Burroughs Welcome Fund Career Award at the Scientific Interface, an Army Research Office Young Investigator Award, and The Johns Hopkins University Whiting School of Engineering start-up funds to C L. Author contributions: S W G designed study, performed experiments, analyzed data, developed template, and wrote the paper; C Y developed robot tail mechanism and assisted with robot experiments; R O and Z R developed robot; C L designed and supervised study, assisted with template development, and wrote the paper. References Kaspari, M , Weiser M D 2016 The Size-Grain Hypothesis and Interspecific Scaling in Ants 13 530 -- 8 Dickinson M H, Farley C T, Full R J, Koehl M A R, Kram R and Lehman S 2000 How Animals Move : An Integrative View 288 100 -- 7 Watson J T, Ritzmann R E, Zill S N and Pollack A J 2002 Control of obstacle climbing in the cockroach, Blaberus discoidalis. I. Kinematics J. Comp. Physiol. A 188 39 -- 53 Blaesing B and Cruse H 2004 Stick insect locomotion in a complex environment: climbing over large gaps J. Exp. Biol. 207 1273 -- 86 Kohlsdorf T and Biewener A A 2006 Negotiating obstacles: Running kinematics of the lizard Sceloporus malachiticus J. Zool. 270 359 -- 71 Sponberg S and Full R J 2008 Neuromechanical response of musculo-skeletal structures in cockroaches during rapid running on rough terrain. J. Exp. Biol. 211 433 -- 46 Jayne B C and Riley M a 2007 Scaling of the axial morphology and gap-bridging ability of the brown tree snake, Boiga irregularis. J. Exp. Biol. 210 1148 -- 60 Li C, Pullin A O, Haldane D W, Lam H K, Fearing R S and Full R J 2015 Terradynamically streamlined shapes in animals and robots enhance traversability through densely cluttered terrain 29 [1] [2] [3] [4] [5] [6] [7] [8] Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu Bioinspir. Biomim 10 [9] Alexander R M 2003 Princliples of animal locomotion (Princeton University Press) Goldman D I, Chen T S, Dudek D M and Full R J 2006 Dynamics of rapid vertical climbing in cockroaches reveals a template. J. Exp. Biol. 209 2990 -- 3000 Theunissen L M and Durr V 2013 Insects use two distinct classes of steps during unrestrained locomotion PLoS One 8 1 -- 18 Higham T E, Korchari P and Mcbrayer L D 2011 How to climb a tree: Lizards accelerate faster, but pause more, when escaping on vertical surfaces Biol. J. Linn. Soc. 102 83 -- 90 Byrnes G and Jayne B C 2014 Gripping during climbing of arboreal snakes may be safe but not economical. Biol. Lett. 1020140434- Byrnes G and Jayne B C 2010 Substrate diameter and compliance affect the gripping strategies and locomotor mode of climbing boa constrictors. J. Exp. Biol. 213 4249 -- 56 Autumn K, Hsieh S T, Dudek D M, Chen J, Chitaphan C and Full R J 2006 Dynamics of geckos running vertically. J. Exp. Biol. 209 260 -- 72 Ritzmann R E, Pollack A J, Archinal J, Ridgel A L and Quinn R D 2005 Descending control of body attitude in the cockroach Blaberus discoidalis and its role in incline climbing J. Comp. Physiol. A 191 253 -- 64 Mongeau J, Mcrae B, Jusufi A, Birkmeyer P, Hoover A M, Fearing R and Full R J 2012 Rapid Inversion : Running Animals and Robots Swing like a Pendulum under Ledges PLoS One 7 Cruse H 1976 The Control of Body Position in the Stick Insect (Carausius morosus), when Walking over Uneven Surfaces Biol. Cybern. 33 25 -- 33 Byrnes G and Jayne B C 2012 The effects of three-dimensional gap orientation on bridging performance and behavior of brown tree snakes (Boiga irregularis) J. Exp. Biol. 215 2611 -- 20 Hoefer K M and Jayne B C 2013 Three-Dimensional Locations of Destinations Have Species- Dependent Effects on the Choice of Paths and the Gap-Bridging Performance of Arboreal Snakes J. Exp. Zool. Part A 319 124 -- 37 Libby T, Moore T Y, Chang-Siu E, Li D, Cohen D J, Jusufi A and Full R J 2012 Tail-assisted pitch control in lizards, robots and dinosaurs Nature 481 181 -- 4 Watson J T, Ritzmann R E and Pollack A J 2002 Control of climbing behavior in the cockroach, Blaberus discoidalis. II. Motor activities associated with joint movement J. Comp. Physiol. A 188 55 -- 69 Harley C M, English B a and Ritzmann R E 2009 Characterization of obstacle negotiation behaviors in the cockroach, Blaberus discoidalis. J. Exp. Biol. 212 1463 -- 76 Olberding J P, McBrayer L D and Higham T E 2012 Performance and three-dimensional kinematics of bipedal lizards during obstacle negotiation J. Exp. Biol. 215 247 -- 55 Parker S E and McBrayer L D 2016 The effects of multiple obstacles on the locomotor behavior and performance of a terrestrial lizard J. Exp. Biol. 219 1004 -- 13 Zurek D B and Gilbert C 2014 Static antennae act as locomotory guides that compensate for visual motion blur in a diurnal , keen-eyed predator Proc. R. Soc. B 281 20133072 30 [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] [29] [30] [31] [32] [33] [34] [35] [36] [37] [38] [39] [40] [41] [42] [43] [44] Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu Birn-Jeffery a. V. and Daley M 2012 Birds achieve high robustness in uneven terrain through active control of landing conditions J. Exp. Biol. 215 2117 -- 27 Spagna J C, Goldman D I, Lin P-C, Koditschek D E and Full R J 2007 Distributed mechanical feedback in arthropods and robots simplifies control of rapid running on challenging terrain. Bioinspir. Biomim. 2 9 -- 18 Daley M and Biewener A A 2006 Running over rough terrain reveals limb control for intrinsic stability. Proc. Natl. Acad. Sci. U. S. A. 103 15681 -- 6 Collins C E, Self J D, Anderson R A and McBrayer L D 2013 Rock-dwelling lizards exhibit less sensitivity of sprint speed to increases in substrate rugosity Zoology 116 151 -- 8 Jayaram K and Full R J 2016 Cockroaches traverse crevices, crawl rapidly in confined spaces, and inspire a soft, legged robot Proc. Natl. Acad. Sci. 113 950 -- 7 Altendorfer R, Moore N, Komsuoglu H, Buehler M, Brown H B, Mcmordie D, Saranli U, Full R and Koditschek D E 2001 RHex: A biologically inspired hexapod runner Auton. Robots 11 207 -- 13 Brill A L, De A, Johnson A M and Koditschek D E 2015 Tail-assisted rigid and compliant legged leaping IEEE Int. Conf. Intell. Robot. Syst. 6304 -- 11 Johnson A M and Koditschek D E 2013 Toward a vocabulary of legged leaping Proc. - IEEE Int. Conf. Robot. Autom. 2568 -- 75 Haynes G C, Pusey J, Knopf R, Johnson A and Koditschek D 2012 Laboratory on legs: an architecture for adjustable morphology with legged robots SPIE Defense, Secur. Sens. 838783870 -- 83870W Chou Y C, Huang K J, Yu W S and Lin P C 2015 Model-based development of leaping in a hexapod robot IEEE Trans. Robot. 31 40 -- 54 Full R J and Koditschek D E 1999 Templates and anchors: neuromechanical hypotheses of legged locomotion on land J. Exp. Biol. 2 3 -- 125 Dürr V 2001 Stereotypic leg searching movements in the stick insect: kinematic analysis, behavioural context and simulation. J. Exp. Biol. 204 1589 -- 604 Cowan N J and Full R J 2006 Task-level control of rapid wall following in the American cockroach J. Exp. Biol. 209 3043 -- 3043 Okada J and Toh Y 2004 Spatio-temporal patterns of antennal movements in the searching cockroach. J. Exp. Biol. 207 3693 -- 706 Dean J and Wendler G 1983 Stick Insect Locomotion on a Walking Wheel: Interleg Coordination of Leg Position J. Exp. Biol. 103 75 -- 94 Theunissen L M, Vikram S and Dürr V 2014 Spatial coordination of foot contacts in unrestrained climbing insects. J. Exp. Biol. 217 3242 -- 53 Lee J, Sponberg S, Loh O, Lamperski A, Full R and Cowan N 2007 Templates and Anchors for Antenna-Based Wall Following in Cockroaches and Robots IEEE Trans Robot 24 1 -- 14 Okada J and Toh Y 2006 Active tactile sensing for localization of objects by the cockroach antenna J. Comp. Physiol. A 192 715 -- 26 [45] Ritzmann R E, Harley C M, Daltorio K A, Tietz B R, Pollack A J, Bender J A, Guo P, 31 [46] [47] [48] [49] [50] [51] [52] [53] [54] [55] [56] [57] [58] [59] [60] [61] Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu Horomanski A L, Kathman N D, Nieuwoudt C, Brown A E and Quinn R D 2012 Deciding which way to go: How do insects alter movements to negotiate barriers Front. Neurosci. 6 1 -- 10 Okada J and Toh Y 2000 The role of antennal hair plates in object-guided tactile orientation of the cockroach (Periplaneta americana) J. Comp. Physiol. A 186 849 -- 57 Daltorio K A, Mirletz B T, Sterenstein A, Cheng J C, Watson A, Kesavan M, Bender J A, Ritzmann R E and Quinn R D 2014 How Cockroaches Employ Wall-Following for Exploration 72 -- 83 Tucker D B and Mcbrayer L D 2012 Overcoming obstacles: The effect of obstacles on locomotor performance and behaviour Biol. J. Linn. Soc. 107 813 -- 23 More H L, Hutchinson J R, Collins D F, Weber D J, Aung S K H and Donelan J M 2010 Scaling of sensorimotor control in terrestrial mammals. Proc. Biol. Sci. 277 3563 -- 8 Jayaram K, Mongeau J, Mcrae B and Full R J 2010 High-speed horizontal to vertical transistions in running cockroaches reveals a priciple of robustness Integr. Comp. Biol. E83 Kubow T M 1999 The role of the mechanical system in control: a hypothesis of self-stabilization in hexapedal runners Philos. Trans. R. Soc. B Biol. Sci. 354 849 -- 61 Jindrich D L and Full R J 2002 Dynamic stabilization of rapid hexapedal locomotion. J. Exp. Biol. 205 2803 -- 23 Dudek D M and Full R J 2006 Passive mechanical properties of legs from running insects. J. Exp. Biol. 209 1502 -- 15 Khatib O 1986 Real-Time Obstacle Avoidance for Manipulators and Mobile Robots Int. J. Rob. Res. 5 90 -- 8 Rimon E and Koditschek D E 1992 Exact Robot Navigation using artificial Potential Functions IEEE Trans. Robot. Autom. 8 501 -- 18 Leonard J J and Durrant-Whyte H F 1991 Simultaneous map building and localization for an autonomous mobile robot Intell. Robot. Syst. 3 1442 -- 7 Koditschek D E 1987 Exact robot navigation by means of potential functions: Some topological considerations IEEE Int. Conf. Robot. Autom. 4 1 -- 6 Thrun S, Burgard W and Fox D 2000 A real-time algorithm for mobile robot mapping with applications to\nmulti-robot and 3D mapping IEEE Int. Conf. Robot. Autom. 1 321 -- 8 Birkmeyer P, Peterson K and Fearing R S 2009 DASH: A dynamic 16g hexapedal robot IEEE/RSJ Int. Conf. Intell. Robot. Syst. 2683 -- 9 Hoover A M, Burden S, Fu X Y, Sastry S S and Fearing R S 2010 Bio-inspired design and dynamic maneuverability of a minimally actuated six-legged robot IEEE/RSJ Int. Conf. Intell. Robot. Syst. 869 -- 76 Lewinger W a., Harley C M, Watson M S, Branicky M S, Ritzmann R E and Quinn R D 2009 Animal-inspired sensing for autonomously climbing or avoiding obstacles Appl. Bionics Biomech. 6 43 -- 61 [62] Schmitt J and Holmes P 2000 Mechanical models for insect locomotion: Dynamics and stability in the horizontal plane I: Theory Biol. Cybern. 83 501 -- 27 32 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu [63] [64] [65] Schmitt J and Holmes P 2000 Mechanical models for insect locomotion: dynamics and stability in the horizontal plane II. Application. Biol. Cybern. 83 517 -- 27 Boggess M J, Schroer R T, Quinn R D and Ritzmann R E 2004 Mechanized cockroach footpaths enable cockroach-like mobility IEEE Int. Conf. Robot. Autom. 3 Schmitt J and Holmes P 2001 Mechanical models for insect locomotion: Stability and parameter studies Phys. D Nonlinear Phenom. 156 139 -- 68 [66] Haldane J B S 1973 On Being the Right Size Annu. Rev. Microbiol. 27 119 -- 32 [67] [68] [69] [70] Gart S and Li C 2017 Dynamic locomotion of insects and legged robots over large 3-D obstacles II. Body-terrain interaction affects high bump traversal Bioinspir. Biomim Marvi H and Hu D L 2012 Friction enhancement in concertina locomotion of snakes J. R. Soc. Interface 9 3067 -- 80 Bell W J, Roth L M and Nalepa C A 2007 Cockroaches. Ecology, Behaviour and Natural History (Baltimore, MD, USA: The Johns Hopkins University Press) Aguilar J, Zhang T, Qian F, Kingsbury M, McInroe B, Mazouchova N, Li C, Maladen R, Gong C, Travers M, Hatton R L, Choset H, Umbanhowar P B and Goldman D I 2016 A review on locomotion robophysics: the study of movement at the intersection of robotics, soft matter and dynamical systems Reports Prog. Phys. 1 -- 61 [71] Full B Y R J and Tu M S 1990 Mechanics of six-legged runners 148 129 -- 46 Qian F, Daffon K, Zhang T and Goldman D I 2013 an Automated System for Systematic Testing of Locomotion on Heterogeneous Granular Media Nature-Inspired Mob. Robot. 547 -- 54 Holmes P, Full R J, Koditschek D and Guckenheimer J 2006 The Dynamics of Legged Locomotion: Models, Analyses, and Challenges SIAM Rev. 48 207 -- 304 Jusufi a, Kawano D T, Libby T and Full R J 2010 Righting and turning in mid-air using appendage inertia: reptile tails, analytical models and bio-inspired robots. Bioinspir. Biomim. 5 45001 Koditschek D E, Full R J and Buehler M 2004 Mechanical aspects of legged locomotion control Arthropod Struct. Dev. 33 251 -- 72 Crall J D, Gravish N, Mountcastle A M and Combes S A 2015 BEEtag : A Low-Cost , Image- Based Tracking System for the Study of Animal Behavior and Locomotion PLoS One 10 e0136487 Ting L H, Blickhan R and Full R J 1994 Dynamic and Static Stability in Hexapedal Runners J. Exp. Biol. 197 251 -- 69 Halloy J, Sempo G, Caprari G, Rivault C, Asadpour M, Tâche F, Saïd I, Durier V, Canonge S, Amé J M, Detrain C, Correll N, Martinoli A, Mondada F, Siegwart R and Deneubourg J L 2007 Social integration of robots into groups of cockroaches to control self-organized choices Science (80-. ). 318 1155 -- 8 Hedrick T L 2008 Software techniques for two- and three-dimensional kinematic measurements of biological and biomimetic systems Bioinspir. Biomim 3 Alexander R M N and Jayes A S 1983 A dynamic similarity hypothesis for the gaits of quadrupedal mammals J. Zool. 201 135 -- 52 33 [72] [73] [74] [75] [76] [77] [78] [79] [80] Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu [81] [82] Seyfarth A, Geyer H, Günther M and Blickhan R 2002 A movement criterion for running J. Biomech. 35 649 -- 55 Seyfarth A 2003 Swing-leg retraction: a simple control model for stable running J. Exp. Biol. 206 2547 -- 55 [83] Blickhan R 1989 The spring-mass model for running and hopping J. Biomech. 22 1217 -- 27 [84] [85] [86] [87] [88] [89] [90] [91] [92] [93] [94] [95] Bekker M G 1960 Off-the-road locomotion; research and development in terramechanics (Ann Arbor, MI: University of Michigan Press) Meirion-Griffith G and Spenko M 2010 An empirical study of the terramechanics of small unmanned ground vehicles IEEE Aerosp. Conf. 1 -- 6 Gray J and Lissmann H W 1950 The kinetics of the locomotion of the grass-snake J. Exp. Biol. 26 354 -- 67 Gonzalez-Gomez J, Zhang H, Boemo E and Zhang J 2006 Locomotion of a Modular Robot with Eight Pitch-Yaw-Connecting Modules 9th Int. Conf. Climbing Walk. Robot. Ohno H and Hirose S 2001 Design of slim slime robot and its gait of locomotion IEEE/RSJ Int. Conf. Intell. Robot. Syst. 2 707 -- 15 Tesch M, Lipkin K, Brown I, Hatton R, Peck A, Rembisz J and Choset H 2009 Parameterized and Scripted Gaits for Modular Snake Robots Adv. Robot. 23 1131 -- 58 Boxerbaum A S, Oro J, Peterson G and Quinn R D 2008 The latest generation Whegs robot features a passive-compliant body joint IEEE/RSJ Int. Conf. Intell. Robot. Syst. 1636 -- 41 Russell P 1975 A contribution to the functional analysis of the foot of the Tokay, Gekko gecko (Reptilia : Gekkonidae) J. Zool. Lond. 176 437 -- 76 Gorb S N 2002 Structural Design and Biomechanics of Friction-Based Releasable Attachment Devices in Insects Integr. Comp. Biol. 42 1127 -- 39 J B W, Roth L M and Nalepa C A 2007 Cockroaches. Ecology, Behaviour and Natural History vol 1 Lillywhite H B, LaFrentz J R, Lin Y C and Tu M C 2000 The Cantilever Abilities of Snakes J. Herpetol. 34 523 -- 8 Li C, Umbanhowar P B, Komsuoglu H, Koditschek D E and Goldman D I 2009 Sensitive dependence of the motion of a legged robot on granular media Proc. Natl. Acad. Sci. 106 9932 -- 9932 [96] Li C, Zhang T and Goldman D I 2013 A Terradynamics of Legged Locomotion on Granular Media Science (80-. ). 339 1408 -- 12 34 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu Supplementary Figures Figure S1. Histogram of approach speed for animal finite gap experiments. Vertical dashed line represents an approach speed of 33 cm s−1 or Fr = 1.5, above which the animal is unlikely able to react to the obstacle in time and likely falls passively (see section 3.1.4). 35 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu Figure S2. Approach speed (a, b) and initial body pitch angular velocity (c, d) is higher during successful traversal than failure for both the animal (a, c) and the robot (b, d). Red filled circles represent successful traversal and blue open circles represent failure to traverse (including grip failure). Error bars represent ± 1 s.d. in (c, d, e, f). Asterisks in indicate a statistically significant difference between the cases of success and failure (P < 0.05, multiple logistic regression). 36 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu Figure S3. Comparison of (a) animal and (b) robot experiments to template predictions of whether each trial results in successful traversal (red) or failure (blue). Circles are experimental data. Red and blue shaded regions are success and failure regimes predicted by the template. 37 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu Figure S4. Dynamic locomotion of the animal and robot over a 0.2 body length gap obstacle. Representative trials of successful traversal (red solid curves) and failure to bridge the gap (blue dashed curves). (a, c) Head height as a function of forward position of the head. (b, d) Body pitch as a function of forward of the head. Solid red and blue dashed curves and shaded areas represent means ± 1 s.d. for the cases of success and failure. For simplicity, only movement perpendicular to the gap (within the x-z plane) is shown. 38 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu Figure S5. Dynamic locomotion of the animal and robot over a 0.4 body length gap obstacle. Representative trials of successful traversal (red solid curves) and failure to bridge the gap (blue dashed curves). (a, c) Head height as a function of forward of the head. (b, d) Body pitch as a function of forward of the head. Solid red and blue dashed curves and shaded areas represent means ± 1 s.d. for the cases of success and failure. For simplicity, only movement perpendicular to the gap (within the x-z plane) is shown. 39 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu Figure S6. Dynamic locomotion of the animal and robot over a 0.6 body length gap obstacle. Representative trials of successful traversal (red solid curves) and failure to bridge the gap (blue dashed curves). (a, c) Head height as a function of forward of the head. (b, d) Body pitch as a function of forward of the head. Solid red and blue dashed curves and shaded areas represent means ± 1 s.d. for the cases of success and failure. For simplicity, only movement perpendicular to the gap (within the x-z plane) is shown. 40 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu Figure S7. Dynamic locomotion of the animal and robot over a 1 body length gap obstacle. Representative trials of successful traversal (red solid curves) and failure to bridge the gap (blue dashed curves). (a, c) Head height as a function of forward of the head. * (b, d) Body pitch as a function of forward of the head. Solid red and blue dashed curves and shaded areas represent means ± 1 s.d. for the cases of success and failure. For simplicity, only movement perpendicular to the gap (within the x-z plane) is shown. * We noted that when the animal failed to traverse, it often rebounded backwards after impacting the far side of the gap, but we could not track this motion due to the body and gap ledge obscuring the tag. Therefore, we truncated the data at the far edge of the gap. 41 Bioinspiration & Biomimetics (2018), 13, 026006; https://li.me.jhu.edu Supplementary Video 1 https://www.youtube.com/watch?v=woS_BzPiqjc Supplementary Video 2 https://www.youtube.com/watch?v=UfM-HNBZizM Supplementary Video 3 https://www.youtube.com/watch?v=yXKVaG9ssiU Supplementary Video 4 https://www.youtube.com/watch?v=wldsMyd_NMA 42
1702.03992
2
1702
2019-07-11T20:08:06
High-throughput nanofluidic device for one-dimensional confined detection of single fluorophores
[ "physics.bio-ph" ]
Ensemble averaging experiments may conceal many fundamental molecular interactions. To overcome that, a high-throughput detection of single molecules or colloidal nanodots is crucial for biomedical, nanoelectronic, and solid-state applications. One-dimensional (1D) discrete flow of nanoscale objects is an efficient approach in this direction. The development of simple and cost-effective nanofluidic devices is a critical step to realise 1D flow. This letter presents a nanofabrication technique using shadow-angle-electron-beam-deposition for a high-throughput preparation of parallel nanofluidic channels. These were used to flow and detect DNA, carbon-nanodots, and organic fluorophores. The 1D molecular mass transport was performed using electro-osmotic flow. The 1D flow behaviour was identified and analysed using two-focus fluorescence correlation spectroscopy (2fFCS). A range of flow velocities of single molecules was achieved. The transitions of single molecules or nanodots through the two foci were quantitatively analysed using confocal scanning imaging, correlative photon detection, and burst size distribution analysis. The results suggest an efficient nanofabrication technique is developed to prepare nanofluidic devices. This first demonstration of high-throughput nanochannel fabrication process and using 2fFCS-based single molecule flow detection should have a potential impact on ultra-sensitive biomedical diagnostics and studying biomolecular interactions as well as nanomaterials.
physics.bio-ph
physics
Single-molecule nanofluidics in a uniform dielectric confinement with resolved molecular shot noise Siddharth Ghosh†,∗ Narain Karedla,† and Ingo Gregor III. Institute of Physics -- Biophysics and Complex Systems, University of Gottingen, Gottingen, Germany Until now, we could not engineer Nature's ability to dynamically handle single molecules in tunnelling-nantubes or pore-forming proteins. Consistent handling of individual single molecules in an extended flow will be of paramount importance to fundamental molecular studies and tech- nological benefits, like single-molecule level separation and sorting for early biomedical diagnostics, microscopic studies of molecular interactions and electron/optical microscopy of molecules and nano- materials. However, at nanometre lenghtscales, dynamics of single molecules can only be reliably resolved if the flow is confined within a uniform dielectric environment as interacting surfaces mod- ify electronic properties of the molecules. This report presents an effective dynamic nanofluidic detection of optically active single molecules. We developed a solid-state method of nanofabricating multiplexed all-silica nanofluidic devices using shadow-angle electron-beam-deposition. In nanoflu- idic flow of single molecules, a uniform dielectric environment is essential to avoid fluctuations in electrodynamic interaction for reproducible photophysical response in dynamic single-molecule fluorescence experiments. The device enabled us to dynamically (electrokinetically) transport indi- vidual single molecules, like biologically relevant fluorescent probes (carbon nanodots and organic fluorophores) and small fragments DNA. We analysed the electrokinetic 1D molecular mass trans- port using two-focus fluorescence correlation spectroscopy (2fFCS) and showed confinement-induced modified molecular interactions while resolving molecular shot noise, which can misrepresent as different size or species of molecules. This first demonstration of high-throughput nanochannel fabrication and 2fFCS-based 1D confined detection of fast moving single molecules opens the av- enue for single-molecule experiments where manipulation of dynamics is necessary, specifically for ultra-sensitive separation of biomolecules. I. INTRODUCTION Due to Brownian motion, it is impossible to keep a molecule within the detection volume for an extended pe- riod of time [1, 2]. Sorting small biomolecules with single- molecule level resolution is still a challenge [3 -- 5] even af- ter 30 years of single-molecule detection [6 -- 8]. Nanoflu- idics [9] can provide means to circumvent these problems for various scientific applications [10 -- 12]. Nanofluidic devices comprising arrays of nanochannels with diame- ters less than 100 nm have become significant for analyt- ical investigation of DNA optical mapping [13 -- 16], sin- gle virus and nanoparticle detection as well as isolation [5, 17 -- 22], ion trapping [23], and energy harvesting [24]. Nanochannels can allow systematic studies of single par- ticles from molecules to viruses over long periods of time by suppressing the thermal motions in two directions -- 1D fluidic confinement. Even today, there is no simple and efficient way to fabricate arrays of enclosed 1D nanofluidic channels. Dif- ferent methods have been published describing the fab- rication processes of such devices.[25 -- 29] However, the ∗ [email protected]; Current Affiliations: Single-Molecule Optics Group, Huygens Laboratory, Leiden Institute of Physics, Lei- den University, The Netherlands.; Centre for Misfolding Diseases, Department of Chemistry and Maxwell Centre, Cavendish Lab- oratory, University of Cambridge, Cambridge, UK. † International Max Planck Research School for Physics of Biolog- ical and Complex Systems, Gottingen, Germany. majority of them are technically challenging, time con- suming, and not easily applicable for high-throughput production. Another problem is photophysical prop- erties of fluophores change due to electrodynamic in- teractions with nonuniform semiconductor-insulator or metal-insulator interfaces.[28, 30 -- 33] Nonuniform dielec- tric nanometre confinements in dynamic experiments of single-molecule fluorescence result to misinterpretation of acquired signal or artefacts. In a uniform dielectric con- finement, molecules can be reliably identified based on their intrinsic photophysics. This is not possible in envi- ronments with nonuniform electrodynamic interactions. Here, we present a simple nanofabrication technique to create all-silica nanochannels using electron beam lithog- raphy (EBL) and shadow-angle electron-beam-deposition (SAEBD), which are suitable for multiplexed dynamic single-molecule detection. The diffraction limited all- silica nanochannels were used to quantify 1D flow and dif- fusion of single molecules, such as small DNA molecules labelled with single organic fluorophores, carbon nan- odots (CND) [34], and organic fluorophores by detecting their fluorescence. Using two-focus fluorescence corre- lation spectroscopy (2fFCS) [35 -- 38], we analysed tran- sitions of single molecules through the nanochannels and quantitatively analysed their flow velocities. The nanofluidic detection of single molecules presented here have been preformed inside a cross-sectional diameters ranging from 30 nm to 100 nm. As far our knowledge goes, dynamic 1D manipulation and 2fFCS investiga- tion of single molecule (of sub-3 nm sizes) inside uniform dielectric nanofluidic confinement has not been demon- 9 1 0 2 l u J 1 1 ] h p - o i b . s c i s y h p [ 2 v 2 9 9 3 0 . 2 0 7 1 : v i X r a strated earlier. II. FABRICATION OF NANOCHANNEL The process steps to create enclosed nanochannels in- volve fabrication of open nanochannels (trenches) using EBL and reactive ion etching (RIE). In a final step, the trenches are closed using SAEBD. SAEBD utilises the ballistic path of the electron beam (e-beam) assisted evaporation.[39] When a collimated beam of evaporated material hits onto an open nano-trench at shallow inci- dent angle, no deposition occurs in the shadowed region. [40, 41] Deposition on the exposed sites causes a growth of material that can enclose a large number of parallel trenches (depending on the beam diameter) leaving the shadowed regions open as paths for fluids. The process is unaffected by nanometre sized residues of the e-beam resists and does not require an atomically clean surface, unlike any wafer bonding-based process.[32] For photo- physical reasons, the nanochannels were prepared using pure silicon dioxide (silica, SiO2). We demonstrate the SAEBD process by scanning electron microscopy (SEM) images of the intermediate steps. Here, we first demon- strate the process using silicon to obtain intermediate steps because silica shows low contrast in SEM due to its dielectric nature. A. Nano-trenches Figure 1a shows a schematic flow-chart of creating nanochannels on Si-[100] wafers. The width of the nano- trenches (i.e. the width of the final nanochannels) can be optimised by the e-beam exposure of the EBL to the positive resist. Different widths of nano-trenches were created ranging from 30 nm to 100 nm. In Fig- ure 1b, 65 nm and 100 nm wide nano-trenches are shown. The lithographed e-beam resist acted as a mask for RIE to etch the final nano-trenches on silicon. We checked the depth of the trenches using AFM. For 40 nm wide trenches we found depths in the range of 35 nm to 40 nm (supplementary Figure 1). B. Nanochannels -- enclosing nano-trenches In the next step, SAEBD was used to enclose the nano-trenches to create closed nanochannels. Figure 1c schematically explains the concept of the SAEBD pro- cess: a high-energy e-beam is bent by a magnet onto the reservoir of material that sublimates and finally de- posits onto the substrate. Figure 1d schematically ex- plains the role of the deposition angle (θ) between the surface of the substrate and the vapour. The angular deposition creates a shadowed region that is unexposed to the depositing material. The deposited and shadowed regions are colour coded in red and yellow, respectively. 2 To demonstrate this, an array of 5 mm long nano-trenches were cross-sectioned using a wafer sawing instrument to observe the intermediate steps while performing SAEBD. Figure 1e schematically represents a high angle deposi- tion which was experimentally performed at 80◦ as shown in Figure 1f. As shown in the time evolution schematic, SAEBD closes the nano-trenches leaving apart a void. By decreasing θ the unexposed area increases. At an acute angle close to 0◦, the growth is nearly parallel to the surface of the substrate as shown in Figure 1e-h. Fig- ure 1g schematically represents a low angle deposition and experimental demonstration of that is shown in Fig- ure 1f at 45◦. Due to instrumental constraints, it was not possible to achieve a deposition at an acute angle close to 0◦. Nevertheless, satisfactory results were obtained using θ = 45◦ as shown in Figure 1i-k. Here, 60 nm ti- tanium was deposited on the open nano-trenches at an angle of 45◦ with a deposition rate of 1A/s at a pressure of 2 × 10−6 mbar. We estimate that at an angle 30◦ a high-quality flat edge will be formed. We used a focused ion beam (FIB) to investigate the cross-sections of the nanochannels. To avoid ion beam induced damage in FIB, the top part of the cutting re- gion was protected with thin metallic layers. We de- posited two thin-films on the top surface of the enclosed nanochannels. Figure 1i - k show SEM images of the milled regions from low to high magnification. In Fig- ure 1j, the first layer (Si-1) is the silicon substrate on which nano-trenches were fabricated. The second layer is titanium (Ti-2) that was deposited under 45◦ SAEBD. Layers Pt-3 and Pt-4 are platinum layers of 100 nm and 450 nm, respectively, which acted as protective layers to avoid FIB induced damage. In Figure 1k, we observe the magnified cross-section of an enclosed nanochannel. As expected, SAEBD growth of titanium produces a well defined flat layer closing the nanochannel. The verti- cal thickness of the titanium film is 47.8 nm, This corre- sponds well with the experimental settings, that were set to deposition of 60 nm Ti under an angle of θ = 45◦ that should result in a layer of 51 nm vertical thickness. After this proof of concept we prepared fused silica based nanofluidic devices. Here, 5 nm of gold film was sputter-coated prior to spin-coating the e-beam resist on the silica wafer. This is necessary to reduce the charging effect of silica under EBL. After RIE, the obtained nano- trenches in the silica wafer were enclosed with SAEBD using silicon dioxide under 45◦. III. RESULTS AND DISCUSSION A. Nanofluidic device and 2fFCS detection scheme The design of the silica based nanofluidic device to perform single molecule experiments is schematically il- lustrated in Figure 2a. Two reservoirs with a diameter of a few millimetres were sand blasted on the silica wafers using 70 µm silica particles prior to the nano-fabrication process. They are separated by a distance of about one millimetre and serve as convenient macroscopic in- and outlet for the fluids. Each of these reservoir is connected to an array of parallel 30 micrometre wide channels. These channels lead to either side of the final array of < 100 nm wide channels. These nanochannels span over a length of l = 200 µm and connect the microchannels. The height of the nano- and the microchannels are the same and are etched in the same step using RIE. In Figure 2a, the white and blue regions indicate the reservoir and the microchannels, respectively (supple- mentary Figs 2, 3, and 4). The red stripes correspond to the array of nanochannels, that are connected to the microchannels. A SEM image of these silica nanochan- nels is shown in the right inset of Figure 2a. For the experiment, we filled the reservoirs with a dilute solution of fluorescent probes in buffer. After filling one reser- voir capillary force transported the fluid through the mi- crochannels and into the nanochannels reservoir. A re- laxation time of 30 s was given to avoid development of trapped air bubbles between two inlets. Then, the sec- ond reservoir was filled. Two 100 µm thick platinum elec- trodes were immersed into the reservoirs (supplementary Figure 5), and an electric field was applied along the nanochannel. This created an electro-osmotic flow [42 -- 45] that uni-directionally transported the fluid through the nanochannels as shown in Figure 2b. To restrict unwanted surface adsorption, we used flu- orescent molecules carrying the same charge as the nanochannels' wall. Pure silica is negatively charged above its isoelectric point (pH(I) = 2).[46, 47] The buffer's pH of 8.5 leads to a considerable amount of nega- tive charges on the walls of silica nanochannels. This con- dition efficiently reduces adsorption of negatively charged molecules in the channels. Figure 2c shows single AlexaFluor 647 molecules (Thermo Fisher, Massachusetts, USA) lined up horizon- tally in all the parallel nanochannels. To immobilise the molecules, the solvent was dried at room temperature leaving behind the molecules in the channels. The pho- ton counts profile shows an average SNR of 90. The im- age was captured using a wide-field optical microscope by exciting the molecules with a 640 nm cw-laser (Coher- ent Laser Systems GmbH, Gottingen, Germany). Beside the evidence of FIB and SEM, this also proves that the nanochannels are properly enclosed and no cross-links between nanochannels are observed, unlike our previous observation. [32] We used 2fFCS to study the dynamics of the molecules inside the nanochannels. The diffraction limited focus of a laser beam in the visible range has a typical diameter of Df ≈ 300 nm to 500 nm. This is much larger than the width of the nanochannels, which is in the range of dnc ≈ 30 nm to 100 nm. Under this condition only a movement along the channel can be detected. Therefore, we can consider the flow inside a nanochannel as quasi 1D. (Figure 2a). In 2fFCS two laser foci are used that are pulsed alternately with a rate of 20 MHz. By perform- 3 ing temporal correlation of the signals from both foci one can accurately determine the times that a fluorescent en- tity takes to move from one focus to the other. In our experiment, the displacement of the foci was ≈ 400 nm and they were carefully aligned with the direction of the nanochannel (Figure 2b). [36] A x-y-z piezo-scanner was used to move the device to the position of the foci and allows to acquire point measurements as well as confocal scan images. B. APD based detection As the first probe of interest, we chose <2 nm sized car- bon nanodots[34] which were later renamed as graphene quantum dots (GQD)[48] and are negatively charged (see supplementary Figure 6). Figure 3a shows a schematic top-view of the device, where the nanochannels appear as dark lines. A confocal scanned image of the x-y plane was recorded, while GQDs were flowing inside the nanochannels (Figure 3b). The applied electric field to induce electro-osmotic flow during the measurement was 15 V/mm. The pixel size of the image is 320 nm with dwell time of 5 ms per pixel. Figure 3c shows a scheme of a y-z section through the nanochannels. In Figure 3d to g we present a series of recorded y-z scanned confocal images. Here, the volume percentage of GQDs relative to the stock solution increases from 1%, 2%, 5%, to 50%, re- spectively. In these scans the pixel size was 100 nm with a dwell time of 2 ms per pixel. In these scans, we observe several periodic point spread functions (PSF) as the flu- orescence is emitted from sub-diffraction sized volumes. At low concentrations of GQDs the fluorescent signals are strongly fluctuating due to the motion of the fluophores out of the excitation focus. As the concentration of GQD increased, the fluctuations of the photon signal decreased and a clear images of the PFSs are obtained. Since the GQDs were close to size of organic fluorophores, it could not have been possible to measure the fast flow in a wide- field emCCD-based setup (supplementary Figure 7 and section emCCD-based detection). C. Electro-osmotic flow of single molecules 1. 2fFCS We investigated the flow velocities of single-molecules in the nanochannnels using 2fFCS. In particular, we used the organic dye Atto 488 (Atto-Tec GmbH, Siegen, Germany) and 48 base-pair dsDNA labelled with Alex- aFluor 647 (IBA GmbH, Gottingen, Germany). The molecules were diluted to a concentration of about 1 nM in aqueous buffer solution (see 'Methods' section). We measured the flow velocities and diffusion of sin- gle molecules inside nanochannels having a width of 30 nm. The applied electric fields range from 27 V/mm to 300 V/mm (see supplementary Figure 9). The temporal cross-correlation functions of the photon traces between the two excitation foci together with the well known dis- tance of the foci allow to accurately determine the flow velocity of the fluorescent probes.[37, 38]. The corre- lated data points from 2fFCS were fitted with the Fokker- Planck equation [49] considering the 1D electro-osmotic flow. A challenge for these measurements is the tight spatial confinement of the molecules not only in the z plane, but also in the y axis. To find the optimal po- sition for the measurements, we recorded a y-z confocal scan around the selected nanochannel. The point that showed the highest number of photons in the scan was then chosen as point of measurements. An exemplary 2fFCS measurement performed at 220 V/mm is shown in Figure 4a. Here, blue and red curves are fits to the two cross-correlation data points. The other two are fits to the autocorrelation data, re- spectively. The velocity and diffusion coefficient from the curve fitting are D = 1.51 × 10−7cm2/s and v = −207µm/s. The negative value of the flow velocity in- fers the direction of the flow from focus 2 to focus 1, which can be altered by changing the applied polarity of the electric field. The linear relationship of electro- osmotic flows at different applied electric fields is plotted in Figure 4b. The linear fit has an r2 value of 0.992. A fluctuation of diffusion coefficient is observed in the range of 1.1 × 10−7 to 9.6 × 10−7 is observed (supplementary Figure 10). However, these values are not very accurate under these conditions, since the driven flow clearly dom- inates the dynamics of the molecules. 2. Photon burst size distribution To confirm that the molecules do not from aggre- gates under the conditions in the confined nanochan- nels we analysed the photon burst sizes of the molecular transits.[50]. Figure 4c shows the burst size distribution (BSD) of the time trace of a flow measurement. The time trace is shown in the inset. The first peak with the low- est photon number is due to the single molecule transits. The second and higher peaks are due to two and more molecules coinciding in the focus. If the molecules move independently as single entities, the BSD is given by a Poissonian distribution. As the fitted curve shows, for the given data this is clearly the case. Figure 4d shows the first 600 ms of the inset 5 ms binned time-trace of Fig- ure 4c. This also shows clearly bursts of single molecule transits. 3. Simulation of confined diffusion For theoretical interpretation, we simulated a similar confined nanofluidic molecular interactions in LAMMPS platform [51] (Figure 4e, also see the supplementary movie S3) where the walls of the nanochannel and particles have same charge. The kymographic analy- 4 sis (Figure 4f-h) of the three different regions inside the nanochannel shows characteristic variations near the walls and at the centre of the wall. Along the centre line of the nanochannel discrete single events are ob- servable along with successive multiple-molecule-events (highlighted in green in Figure 4f). Molecular interac- tions near the wall (Figure 4g) are dominated by a large number of crawling events where molecules were moving along the wall for extended periods of time -- ranging from 10 to 130 frames. These events are highlighted in red. In Figure 4h, at the wall -- fluid-solid interface, both the red and green events are not dominant. Within the diffrac- tion limited confocal detection volumes of 2fFCS, besides the molecular shot noise, the crawling events are another potential cause behind the second less frequent peak in the BSD analysis and variation in diffusion coefficients. Surface chemistry -- a function of surface roughness and ionic contaminants play an important role in electroki- netic transports at nanometre length scale as observed by Golestanian and co-workers [52]. IV. CONCLUSION At the length scale where molecular interactions can be strongly influenced by surface charge, a uniform di- electric confinement removes unwanted artefacts gener- ated by non-uniform electrodynamic interactions. The observed molecular behaviour in the fabricated all-silica nanofluidic channels, demonstrates absence of confine- ment induced artefacts. In summary, this Letter presents an efficient method to fabricate all-silica nanochannels using SAEBD and their usage to reduce thermal motion in two spatial dimensions in single molecule experiments. The nanofabrication process demonstrated here is suit- able for high-throughput and large scale production of nanofluidic chips. SAEBD is not material restricted, un- like the oxidation and bonding based techniques. Large range velocity variations of 1D flow of single organic flu- orophores, sub-2 nm GQDs, and 11 nm DNA molecules were achieved using electro-osmosis inside nanochannels. As far our knowledge goes, this is the first quantitative study of 1D nanofluidic flows in nanochannels at single- molecule level utilising 2fFCS. A broad range of velocities of up to 300 µm/s were achieved by varying the applied electric field. BSD analysis confirms that the observed transits were mainly due to single emitters. All the exper- iments were performed inside nanochannels with a cross- section of 30 nm × 35 nm. The experimental observations are justified with BSD analysis and MD simulation. This simple approach of fabricating nanochannels paves the way towards detecting early onset of any disease at sin- gle molecule level. In future, trapping nanoscale objects of less than 2 nm in size for large residence time should be also feasible using these nanochannels. Biomolecular interactions such as dynamics of DNA, protein aggrega- tion, and structural biology of molecules under physiolog- ical condition can be also studied at single molecule level using the SAEBD based nanofluidic devices. The funda- mental understanding diffusion inside nanofludic channel requires further investigation and opens another path of fundamental research. ACKNOWLEDGEMENT SG thanks the International Max Planck Research School for Physics of Biological and Complex Systems and the Ministry of Science and Culture (Lower Saxony) for awarding an Excellence Stipend/MWK PhD scholar- ship. The work presented here was carried out at the re- search group of Prof Jorg Enderlein. The authors thank the internal funding from the University of Gottingen as- sociated to the Third Institute of Physics for this project. Lastly, the authors are extremely grateful to Prof Ender- lein for providing plenty of scientific advice. Dr Mitja Platen and Ditter Hille has supported us in AFM and building the SAEBD stages, respecitively. Authors also thank Xuxing Lu (Single Molecule Optics Group, Leiden) for suggestions on the simulations. AUTHORS CONTRIBUTION SG has written the manuscript, conceptualised the idea of nanofabrication, prepared the nanofluidic device, per- 5 formed all the measurements and numerical fitting pre- sented in this paper. SG and NK have performed ini- tial measurements together. NK has modified the fitting routine provided by Prof Jorg Enderlein and IG. IG has supervised the project and provided scientific advice. All the authors have read and made adequate corrections in the manuscripts. COMPETING INTERESTS The authors declare no competing interests. ADDITIONAL INFORMATION The supplementary information contains the methods of nanofabrication, figures and design of the nanofludic device, integrating the nanofluidic device to electrodes, carbon nanodots/GQD cation-pi interaction behaviour, EMCCD-based measurements, description of 2fFCS setup, 2fFCS fitting function, control 1D diffusion of single Atto 488 fluorophores, electro-osmotic flow and diffusion of single DNA molecules at different electric fields, and LAMMPS simulation. [1] A. E. Cohen, Physical review letters 91, 235506 (2003). [2] A. E. Cohen and W. Moerner, Proceedings of the Na- [14] S. K. Min, W. Y. Kim, Y. Cho, and K. S. Kim, Nature nanotechnology 6, 162 (2011). tional Academy of Sciences 103, 4362 (2006). [15] B. M. Venkatesan and R. Bashir, Nature nanotechnology [3] P. M. Goodwin, M. E. Johnson, J. C. Martin, W. P. Ambrose, B. L. Marrone, J. H. Jett, and R. A. Keller, Nucleic Acids Research 21, 803 (1993). [4] P. V. Jones, G. L. Salmon, and A. Ros, Analytical chem- istry 89, 1531 (2017). 6, 615 (2011). [16] C. Wang, R. L. Bruce, E. A. Duch, J. V. Patel, J. T. Smith, Y. Astier, B. H. Wunsch, S. Meshram, A. Galan, C. Scerbo, et al., ACS nano 9, 1206 (2015). [17] A. Mitra, B. Deutsch, F. Ignatovich, C. Dykes, and [5] M. J. Skaug, C. Schwemmer, S. Fringes, C. D. Rawlings, L. Novotny, ACS nano 4, 1305 (2010). and A. W. Knoll, Science 359, 1505 (2018). [6] N. J. Dovichi, J. C. Martin, J. H. Jett, M. Trkula, and R. A. Keller, Analytical chemistry 56, 348 (1984). [7] D. C. Nguyen, R. A. Keller, J. H. Jett, and J. C. Martin, Analytical chemistry 59, 2158 (1987). [8] W. E. Moerner and L. Kador, Physical review letters 62, [18] M. N. Hamblin, J. Xuan, D. Maynes, H. D. Tolley, D. M. Belnap, A. T. Woolley, M. L. Lee, and A. R. Hawkins, Lab on a Chip 10, 173 (2010). [19] J.-L. Fraikin, T. Teesalu, C. M. McKenney, E. Ruoslahti, and A. N. Cleland, Nature nanotechnology 6, 308 (2011). [20] A. de la Escosura-Muniz and A. Merko¸ci, Small 7, 675 2535 (1989). (2011). [9] J. C. Eijkel and A. Van Den Berg, Microfluidics and Nanofluidics 1, 249 (2005). [10] J. K. Holt, H. G. Park, Y. Wang, M. Stadermann, A. B. Artyukhin, C. P. Grigoropoulos, A. Noy, and O. Bakajin, Science 312, 1034 (2006). [11] P. Goswami and S. Chakraborty, Langmuir 26, 581 (2009). [21] S. Faez, Y. Lahini, S. Weidlich, R. F. Garmann, K. Won- draczek, M. Zeisberger, M. A. Schmidt, M. Orrit, and V. N. Manoharan, ACS nano 9, 12349 (2015). [22] B. H. Wunsch, J. T. Smith, S. M. Gifford, C. Wang, M. Brink, R. L. Bruce, R. H. Austin, G. Stolovitzky, and Y. Astier, Nature nanotechnology 11, 936 (2016). [23] N. Kavokine, S. Marbach, A. Siria, and L. Bocquet, [12] R. H. Tunuguntla, R. Y. Henley, Y.-C. Yao, T. A. Pham, Nature nanotechnology , 1 (2019). M. Wanunu, and A. Noy, Science 357, 792 (2017). [24] A. Siria, M.-L. Bocquet, and L. Bocquet, Nature Re- [13] F. Persson and J. O. Tegenfeldt, Chemical Society Re- views Chemistry 1, 0091 (2017). views 39, 985 (2010). [25] F. Persson, L. H. Thamdrup, M. B. L. Mikkelsen, S. Jaarlgard, P. Skafte-Pedersen, H. Bruus, and A. Kris- 6 tensen, Nanotechnology 18, 245301 (2007). [26] D. C. Bien, H. W. Lee, and R. M. Saman, ECS Solid State Letters 1, P45 (2012). [27] S. H. Kim, Y. Cui, M. J. Lee, S.-W. Nam, D. Oh, S. H. Kang, Y. S. Kim, and S. Park, Lab on a Chip 11, 348 (2011). [28] F. Westerlund, F. Persson, A. Kristensen, Tegenfeldt, Lab on a Chip 10, 2049 (2010). and J. O. [29] P. Mao and J. Han, Lab on a Chip 5, 837 (2005). [30] R. Hildner, D. Brinks, J. B. Nieder, R. J. Cogdell, and N. F. van Hulst, Science 340, 1448 (2013). [31] E. Wientjes, J. Renger, A. G. Curto, R. Cogdell, and N. F. Van Hulst, Nature communications 5, 4236 (2014). [32] S. Ghosh, M. Kumbhakar, M. Platen, I. Gregor, and J. Enderlein, in Single Molecule Spectroscopy and Super- resolution Imaging VII, Vol. 8950 (International Society for Optics and Photonics, 2014) p. 895008. [33] H. Cheng, Probing Molecular Stoichiometry by Photon Antibunching and Nanofluidics Assisted Imaging in Solu- tion, Ph.D. thesis, Georg-August-Universitat Gottingen (2017). [34] S. Ghosh, A. M. Chizhik, N. Karedla, M. O. Dekaliuk, I. Gregor, H. Schuhmann, M. Seibt, K. Bodensiek, I. A. Schaap, O. Schulz, et al., Nano letters 14, 5656 (2014). [35] M. Brinkmeier, K. Dorre, J. Stephan, and M. Eigen, Analytical chemistry 71, 609 (1999). [36] T. Dertinger, V. Pacheco, I. von der Hocht, R. Hart- mann, I. Gregor, and J. Enderlein, ChemPhysChem 8, 433 (2007). [37] T. Dertinger, A. Loman, B. Ewers, C. B. Muller, B. Kramer, and J. Enderlein, Optics express 16, 14353 (2008). [38] T. J. Arbour and J. Enderlein, Lab on a Chip 10, 1286 (2010). [39] J. Merkel, T. Sontheimer, B. Rech, and C. Becker, Jour- nal of Crystal Growth 367, 126 (2013). [40] S. Egger, A. Ilie, S. Machida, and T. Nakayama, Nano letters 7, 3399 (2007). [41] D. M. Mattox, Handbook of physical vapor deposition (PVD) processing (William Andrew, 2010). [42] X. Huang, M. J. Gordon, and R. N. Zare, Analytical chemistry 60, 1837 (1988). [43] N. A. Patankar and H. H. Hu, Analytical Chemistry 70, 1870 (1998). [44] N. J. Dovichi and J. Zhang, Angewandte Chemie Inter- national Edition 39, 4463 (2000). [45] S. Chakraborty and A. K. Srivastava, Langmuir 23, 12421 (2007). [46] M. Kosmulski, Chemical properties of material surfaces (CRC press, 2001). [47] M. Keswani, S. Raghavan, P. Deymier, and S. Verhaver- beke, Microelectronic Engineering 86, 132 (2009). [48] S. Ghosh, M. Awasthi, M. Ghosh, M. Seibt, and T. A. Niehaus, 2D Materials 3, 041008 (2016). [49] H. Risken, in The Fokker-Planck Equation (Springer, 1996) pp. 63 -- 95. [50] J. Enderlein, D. L. Robbins, W. P. Ambrose, and R. A. Keller, The Journal of Physical Chemistry A 102, 6089 (1998). [51] S. Plimpton, Journal of computational physics 117, 1 (1995). [52] S. Ebbens, D. Gregory, G. Dunderdale, J. Howse, Y. Ibrahim, T. Liverpool, and R. Golestanian, EPL (Eu- rophysics Letters) 106, 58003 (2014). 7 FIG. 1. Nanochannel fabrication using SAEBD. a. Fabrication of nano-trenches on silicon with EBL and RIE. b. SEM of silicon nano-trenches with 62 nm and 100 nm width. c. SAEBD at angle θ. d. Shadows of electron beam. Arrows indicate angular e-beam evaporation. e. Schematic of high angle deposition and f. SEM of deposition at 80◦. g. Schematic of low angle deposition and h. SEM of deposition at 45◦. i. FIB cross-section of the enclosed nanochannels. Two layers of platinum were used to protect the nanochannel form high energy ions. j. Magnified view of the enclosed nanochannel, where Si-1 is the silicon substrate on which nano-trenches were fabricated, Ti-2 is the 50 nm thick titanium layer which was used in SAEBD, Pt-3 and Pt-4 are platinum deposited inside FIB. k. Further magnified view of the nanochannel where region 1 is silicon (Si-1) and region 2 is titanium (Ti-2). angle of deposition =material deposited at anglemagnetelectron beamnon-conformal material deposition65 nmSi-1Ti-2Pt-3Pt-412Enclosed nanochannel400 nm3 μmfeSilicon(100)100nmPMMAe-beamlithographyDevelopedwithMIBK:IPAe-beam exposed PMMARIE for Silicon etchingSilicon nano-trenchhikjg48 nmt = 0cdabNano-trench fabricationShadow angle e-beam deposition (SAEBD) 8 device in a 2. Nanofluidic FIG. 2fFCS setup. a. Schematic top view of nanofluidic devices with an SEM image of real nanochannels (scale bar is 30 µm). b. Side view schematic of the complete experimental set-up where electric field is applied through two reservoirs along the nanochannels using platinum electrodes. Two foci (of 2fFCS) were aligned with nanochannel using a 100× 1.49 NA oil immersion objective lens. Two different linearly polarised pulsed interleaved lasers were used for in two-focus excitation. The emission from flowing single molecules was detected by two APDs. c. A wide-field of image frame showing the presence of single molecule (scale bar is 8 µm). 0.01.00250500Pixels60120180Counts100x/ Oil1.49hνAPD 2NPPulsed Laser 1Pulsed Laser 2DM NP - Normaski prismDM - Dicroic mirrorPH - PinholeL - LensBS - Beam splitterPIE - Pulsed interleaved excitationAPD 1PIEExEyEyExdDyzyxzfD >>dncncXZYpiezo scannerfbPHBScBSLLPt electrodePt electrodea 9 1D Flow of GQDs. a. FIG. 3. Schematic top-view of nanochannels along which an (y-x ) scan was performed. b. Confocal scan image of nanochannels filled with GQDs. c. Schematic cross-sectional side view of nanochannels along which (y-z ) scans were performed. The dashed arrow repre- sent the optical excitation path -- direction from immersion oil of objective lens to nanofluidic device. d, e, and f y-z scan images of nanochannels with an increasing order of GQDs' concentration flowing through nanochannels. g. High concen- tration of GQD. All the horizontal scale bars denotes 2µm. The vertical scale bar denotes photon counts as 0.0 (lowest) to 1.0 (highest). FIG. 4. 1D Flow of 48bp DNA. a. 2fFCS measured correlation plots of photon counts from two APDs. Here, applied electric field was of 220 V/mm. The data was fitted with 1D Fokker-Planck equation. From the fitting, we found diffussion coefficient of 1.51 × 10−7 cm2/s. b. Plot of velocity versus applied voltage along the nanochannel. The data was fitted with linear fit. c. Burst size distribution of single molecule transits fitted with two Poissonian distributions. Total time trace binned with 5 ms is shown in the inset -- the dashed bordered grey region is shown in d. d. A part of the time trace plots of single photon bursts due single DNA molecules transits through focus. MD simulation of 1D diffusion. e. A frame from the simulation of diffusing particles inside a confined channel. Kymograph f. along the centre of the channel (yellow line) shows presence of molecular shot noise (green highlights), g. along the orange dashed line i.e. close to the wall with large degree of wall interaction induced drag or crawling (red highlights), and h. along the blue dashed line at the wall-particle interface with less degree of wall interaction in contrast to g. 0.01.00.5abcdezyfgyxV = 220 Vcorrelation / (hν)2/s2−2261014102103104105time[μs]v= −206.93µm/s0501001502002503000100200300Applied voltage (V)Velocity (µm/s)r2=0.99275 125 175 225 275 27 1890Photons countsFrequency3020100time (ms)0 100 200 300 400 500Photons counts (KHz)dcabeg0 354 708# pixels# frames5002500hf# frames5002500# frames500250030201004050time (ms)x1030 2 4 6 8 10 12Photons counts (KHz)